You are on page 1of 41

Hypersonic Aerodynamics

19EAE442
Unit – I
Basics of Hypersonic Aerodynamics
Syllabus
Basics of Hypersonic Aerodynamics: Introduction to
hypersonic aerodynamics, thin shock layers, entropy layers,
low density and high density flows, hypersonic flight paths,
hypersonic flight similarity parameters, shock wave and
expansion wave relations of inviscid hypersonic flows.
History of Hypersonic Flight
• V-2/WAC corporal lift – off
• It is the first meaning attempt to demonstrate the
use of a multi-stage rocket for achieving high
velocities and high altitudes.
• Rocket is V-2, mounted on the top of V-2 is a
slender, needle like rocket called the WAC corporal
which serves as second stage to the V-2.
• Pen plotters track the V-2 to an altitude of 161km at
a velocity of 1560 m/s, at which the WAC corporal
ignited.
• The slender upper stage accelerates to a maximum
velocity of 2300 m/s and reaches an altitude of 390
km.
• After reaching this peak, the WAC Corporal noses
over and careers back into the atmosphere at over
2200 m/s.
• In so doing, it becomes the first object of human
origin to achieve hypersonic flight—the first time
that any vehicle has flown faster than five times the
speed of sound.
24th Feb, 1949, New Mexico
History of Hypersonic Flight Contd..
• The world’s first spaceship, Vostok (East), with a man on
board was launched into orbit from the Soviet Union on
April 12, 1961.
• The pilot space-navigator of the satellite-spaceship
Vostok is a citizen of the U.S.S.R., Flight Major Yuri
Gagarin.
• The launching of the multistage space rocket was
successful and, after attaining the first escape velocity
and the separation of the last stage of the carrier rocket,
the spaceship went into free flight on around-the-earth
orbit.
• According to preliminary data, the period of the
revolution of the satellite spaceship around the earth is
89.1 min.
• The spaceship with the navigator weighs 4725 kg
(10,418.6 lb), excluding the weight of the final stage of
the carrier rocket.
• Major Gagarin’s orbital craft, called Vostok I, is slowed
by the firing of a retrorocket and enters the atmosphere at
a speed in excess of 25 times the speed of sound
• Moreover, on that day, 12 April 1961, Yuri Gagarin
becomes the first human being in history to experience
hypersonic flight.
History of Hypersonic Flight Contd..
• On 5 May, Alan B. Shepard becomes the second man in space by virtue of a suborbital flight over the
Atlantic Ocean, reaching an altitude of 186.2 km, and entering the atmosphere at a speed above Mach 5.

• On 23 June, U.S. Air Force test pilot Major Robert White flies the X-15 airplane at Mach 5.3, the first
X-15 flight to exceed Mach 5.
Introduction
• Hypersonic flight has special traits, some of which are seen in every hypersonic flight.
• Presence of these particular features during a flight is highly dependent on type of trajectory, configuration
etc.
• In short it is the mission requirement which decides the nature of hypersonic atmosphere encountered by the
flight vehicle.
• Some missions are designed for high deceleration in outer atmosphere during reentry.
• Hence, those flight vehicles experience longer flight duration at high angle of attacks due to which blunt
nosed configuration are generally preferred for such aircrafts.
• On the contrary, some missions are centered on low flight duration with major deceleration closer to earth
surface hence these vehicles have sharp nose and low angle of attack flights.
• Reentry flight path of hypersonic vehicle is thus governed by the parameters called as ballistic parameter and
lifting parameter.
• These parameters are obtained by applying momentum conservation equation in the direction of the flight
path and normal to it.
• Velocity-altitude map of the flight is thus made from the knowledge of these governing flight parameters,
weight and surface area.
• Ballistic parameter is considered for non lifting reentry flights like flight path of Apollo capsule, however
lifting parameter is considered for lifting reentry trajectories like that of space shuttle.
Introduction
• Therefore hypersonic flight vehicles are classified in four different types based on the design constraints
imposed from mission specifications

• Reentry Vehicles (RV): These vehicles are typically launched using rocket propulsion system. Reentry of
these vehicles is controlled by control surfaces. Large angle of attack flight of blunt nosed configurations is
the need of these flights. Space shuttle ( US ), BURAN (Russian), HOPE ( Japan ) and HERMES (European)
are some examples of these kind vehicles.

• Cruise and Acceleration Vehicle (CAV): Slender configurations with low angle of attack flights are main
features of these flights. These vehicles are prepared for high heating loads with ablative cooling system. Air
breathing propulsion system of ram or scramjet type is generally preferred for these vehicles. Sanger, which is
a two stage (TSTO) hypersonic vehicle, has first stage with air breathing propulsion and second stage is
propelled with rocket. Hence first stage of Sanger falls in CAV category for which separation takes place at
Mach 7.
Introduction
• Ascent and Reentry Vehicles (ARV): These vehicles have opposing requirements of their design due to dual
duty of ascent, which is dominated by fuel requirements, and reentry by aero-braking. Rocket or air breathing
propulsion systems can be preferred for these flights. NASP or National Aerospace Plane of US, Space Plane
by Japan and HOTOL are some examples of these vehicles.

• Aeroassisted Orbit Transfer Vehicle (AOTV): This is one more class in which hypersonic vehicles are
classified. Ionisation and hence presence of plasma in the vicinity of the spacecraft is the major concern of
these vehicles.
Hypersonic Flow – What is it?
• There is a conventional rule of thumb that defines hypersonic aerodynamics as
those flows where the Mach number M is greater than 5.

• Hypersonic flow is best defined as that regime where certain physical flow
phenomena become progressively more important as the Mach number is
increased to higher values.

• In some cases, one or more of these phenomena might become important above
Mach 3, whereas in other cases they may not be compelling until Mach 7 or
higher.
Physical characteristics of hypersonic flow
Thin Shock Layers
• From oblique shock theory that, for a given flow deflection angle, the density increase across the
shock wave becomes progressively larger as the Mach number is increased.
• For flow over a hypersonic body, this means that the distance between the body and the shock
wave can be small.
• The flow field between the shock wave and the body is defined as the shock layer, and for
hypersonic speeds this shock layer can be quite thin.
• It is a basic characteristic of hypersonic flows that
shock waves lie close to the body and that the
shock layer is thin.
• This interpretation of shock layer thinness for
calorically perfect gas is also applicable for thermally perfect gas and chemically reacting flow.
• However, complexity of flow field increases due to thin shock layer where the boundary layer
thickness and shock layer thickness become comparable.
Thin Shock Layers
• In turn, this can create some physical complications,
such as the merging of the shock wave itself with a thick, viscous boundary layer growing from the
body surface—a problem that becomes important at low Reynolds numbers.
• At high Reynolds numbers, where the shock layer is essentially inviscid, its thinness can be used to
theoretical advantage, leading to a general analytical approach called thin shock-layer theory.
• A thin shock layer approaches the fluid-dynamic model postulated by Issac Newton in 1687; such
Newtonian theory is simple and straightforward and is frequently used in hypersonic aerodynamics
for approximate calculations
Entropy Layer
• At hypersonic Mach numbers, the shock layer over
the blunt nose is also very thin, with a small
shock-detachment distance d.
• In the nose region, the shock wave is highly curved.
• Entropy of flow increases across a shock wave.
• The stronger the shock, the larger the entropy
increase.
• A streamline passing through the strong, nearly
normal portion of the curved shock near the
centerline of the flow will experience a larger
entropy increase than a neighboring streamline, which passes through a weaker portion of the
shock away from the centerline.
• Therefore, a layer of entropy variation getting formed downstream of the shock is termed as
entropy layer
Entropy Layer
• Hence, there are stronger entropy gradients generated in the nose region; this entropy layer flows
downstream and essentially wets the body for larger distances from the nose.

• Boundary layer along the surface grows inside this entropy layer and is affected by it.

• Because the entropy layer is also a region of strong vorticity, this interaction is sometimes called a
vorticity interaction.
Viscous Interaction
• Consider a boundary layer on a flat plate in a
hypersonic flow, as sketched in Fig.
• A high-velocity, hypersonic flow contains a
large amount of kinetic energy; when this
flow is slowed by viscous effects within the
boundary layer, the lost kinetic energy is
transformed (in part) into internal energy of
the gas— this is called viscous dissipation.
• In turn, the temperature increases within the boundary layer; a typical temperature profile within
the boundary layer is also sketched
Viscous Interaction
• The characteristics of hypersonic boundary layers are dominated by such temperature increases.
• For example, the viscosity coefficient increases with temperature, and this by itself will make the
boundary layer thicker.
• In addition, because the pressure p is constant in the normal direction through a boundary layer, the
increase in temperature T results in a decrease in density ρ through the equation of state ρ = p/RT,
where R is the specific gas constant.
• To pass the required mass flow through the boundary layer at reduced density, the boundary-layer
thickness must be larger.
• Both of these phenomena combine to make hypersonic boundary layers grow more rapidly than at
slower speeds.
Viscous Interaction
• The flat-plate compressible laminar boundary-layer thickness δ grows essentially as
2
𝑀∞
𝛿∝
𝑅𝑒𝑥
• δ varies as the square of M∞, it can become inordinately large at hypersonic speeds.
• The thick boundary layer in hypersonic flow can exert a major displacement effect on the inviscid
flow outside the boundary layer, causing a given body shape to appear much thicker than it really
is.
• Because of the extreme thickness of the boundary-layer flow, the outer inviscid flow is greatly
changed; the changes in the inviscid flow in turn affect the growth of the boundary layer.
• This major interaction between the boundary layer and the outer inviscid flow is called viscous
interaction.
• Viscous interactions can have important effects on the surface-pressure distribution, hence lift,
drag, and stability on hypersonic vehicles.
• Moreover, skin friction and heat transfer are increased by viscous interaction.
• The boundary layer on a hypersonic vehicle can become so thick that it essentially merges with
the shock wave—a merged shock layer.
• When this happens, the shock layer must be treated as fully viscous, and the conventional
boundary layer analysis must be completely abandoned.
High-Temperature Flows
• The kinetic energy of a high-speed, hypersonic flow is dissipated by the influence of friction
within a boundary layer.

• The extreme viscous dissipation that occurs within hypersonic boundary layers can create very
high temperatures—high enough to excite vibrational energy internally within molecules and to
cause dissociation and even ionization within the gas.

• If the surface of a hypersonic vehicle is protected by an ablative heat shield, the products of
ablation are also present in the boundary layer, giving rise to complex hydrocarbon chemical
reactions. On both accounts, we see that the surface of a hypersonic vehicle can be wetted by a
chemically reacting boundary layer.
High-Temperature Flows
• Consider the nose region of a blunt body, as sketched in Fig.
The bow shock wave is normal, or nearly normal, in the
nose region, and the gas temperature behind this strong
shock wave can be enormous at hypersonic speeds.
• For example, Fig. is a plot of temperature behind a normal
shock wave as a function of freestream velocity, for a vehicle
flying at a standard altitude of 52 km.
• Two curves are shown:
• 1) the upper curve, which assumes a calorically perfect
nonreacting gas with the ratio of specific heats γ = 1.4, and
which gives an unrealistically high value of temperature;
• 2) the lower curve, which assumes an equilibrium chemically reacting gas and which is usually
closer to the actual situation.
High-Temperature Flows
• This figure illustrates two points:
• 1) By any account, the temperature in the nose region of a hypersonic vehicle can be extremely
high, for example, reaching approximately 11,000 K at a Mach number of 36 (Apollo reentry).
• 2) The proper inclusion of chemically reacting effects is
vital to the calculation of an accurate shock-layer
temperature; the assumption that g is constant and
equal to 1.4 is no longer valid.
• For a hypersonic flow, not only can the boundary layer
be chemically reacting, but the entire shock layer can be
dominated by chemically reacting flow.
High-Temperature Flows
• When the gas temperature is increased to high values, the gas behaves in a “nonideal” fashion,
specifically as follows:
1) The vibrational energy of the molecules becomes excited, and this causes the specific heats cp
and cv to become functions of temperature. In turn, the ratio of specific heats, γ = cp/cv, also
becomes a function of temperature. For air, this effect becomes important above a temperature of
800 K.
2) As the gas temperature is further increased, chemical reactions can occur. For an equilibrium
chemically reacting gas, cp and cv are functions of both temperature and pressure, and hence γ = f(T,
p).
For air at 1 atm pressure, O2 dissociation (O2  2O) begins at about 2000 K, and the molecular
oxygen is essentially totally dissociated at 4000 K.
At this temperature N2 dissociation (N2  2N) begins and is essentially totally dissociated at 9000 K.
Above a temperature of 9000 K, ions are formed (N N+ + e-, and O O+ + e-), and the gas
becomes a partially ionized plasma.
• All of these phenomena are called high-temperature effects.
High-Temperature Flows
• If the vibrational excitation and chemical reactions take place very rapidly in comparison to the
time it takes for a fluid element to move through the flowfield  vibrational and chemical
equilibrium flow.
• If the opposite is true  non-equilibrium flow.
• High-temperature chemically reacting flows can have an influence on lift, drag, and moments on a
hypersonic vehicle.
• The most dominant aspect of high temperatures in hypersonics is the resultant high heat-transfer
rates to the surface.
• Aerodynamic heating dominates the design of all hypersonic machinery, whether it be a flight
vehicle, a ramjet engine to power such a vehicle, or a wind tunnel to test the vehicle.
• This aerodynamic heating takes the form of heat transfer from the hot boundary layer to the
cooler surface—called convective heating.
• Moreover, if the shock-layer temperature is high enough the thermal radiation emitted by the gas
itself can become important, giving rise to a radiative flux to the surface—called radiative heating.
High-Temperature Flows
• Another consequence of high-temperature flow over hypersonic vehicles is the “communications
blackout” experienced at certain altitudes and velocities during atmospheric entry, where it is
impossible to transmit radio waves either to or from the vehicle.

• This is caused by ionization in the chemically reacting flow, producing free electrons that absorb
radio-frequency radiation.

• Therefore, the accurate prediction of electron density within the flowfield is important.
Low Density Flow
• The average distance the molecule moves between successive collisions is defined as mean free
path, denoted by λ.
• At standard sea-level conditions, for air, λ = 2.176 x 10-7 ft.
• This implies that, at sea level, when you wave your hand through the air the gas itself “feels” like
a continuous medium—a so-called continuum.
• Most aerodynamic problems (more than 99.9%of all applications) are properly addressed by
assuming a continuous medium.
• At an altitude of 342,000 ft, where the air density is much lower, and consequently the mean free
path is much larger than at sea level; indeed, at 342,000 ft, λ =1 ft.
• When you wave your hand through the air, you are more able to feel individual molecular
impacts; the air no longer feels like a continuous substance, but rather like an open region
punctuated by individual, widely spaced particles of matter.
• Under these conditions, the aerodynamic concepts, equations, and results based on the
assumption of a continuum begin to break down.
• When this happens, we have to approach aerodynamics from a different point of view, using
concepts from kinetic theory.
• This regime of aerodynamics is called low-density flow.
Low Density Flow
• For any given flight vehicle, as the altitude progressively increases (hence the density decreases
and λ increases), the assumption of a continuum flow becomes tenuous.
• An altitude can be reached where the conventional viscous flow no-slip conditions begin to fail.
• Specifically, at low densities the flow velocity at the surface, which is normally assumed to be zero
because of friction, takes on a finite value. This is called the velocity-slip condition.
• In analogous fashion, the gas temperature at the surface, which is normally taken as equal to the
surface temperature of the material, now becomes something different. This is called the
temperature slip condition.
• At the onset of these slip effects, the governing equations of the flow are still assumed to be the
familiar continuum-flow equations, except with the proper velocity and temperature-slip
conditions utilized as boundary conditions.
• Finally, the air density can become low enough that only a few molecules impact the surface per
unit time, and after these molecules reflect from the surface they do not interact with the
incoming molecules. This is the regime of free molecule flow.
Low Density Flow
• For the space shuttle, the free molecular regime begins about 150 km (500,000 ft).
• A hypersonic vehicle moving from a very rarified atmosphere to a denser atmosphere will shift
from the free molecular regime, where individual molecular impacts on the surface are
important, to the transition regime, where slip effects are important, and then to the continuum
regime.
• The similarity parameter that governs these different regimes is the Knudsen number, defined as
Kn = λ/L, where L is a characteristic dimension of the body.
• The region where the continuum Navier–Stokes equations hold is described by Kn < 0.2.
• However, slip effects must be included in these equations when Kn > 0.03.
• The effects of free molecular flow begin around a value or Kn = 1 and extend out to the limit of Kn
becoming infinite.
• Hence, the transitional regime is essentially contained within 0.03 < Kn < 1.0.
• If Kn is very small, we have continuum flow; if Kn is very large, we have free molecular flow, and
so forth.
Hypersonic Flight Paths: Velocity-Altitude Maps
• Consider a vehicle flying at a velocity V along a flight path inclined at the angle θ below the local horizontal,
as shown in Fig.
• The forces acting on the vehicle are lift L, drag D, and
weight W; the thrust is assumed to be zero;
hence, we are considering a hypersonic glide vehicle.
• Summing forces along and perpendicular to the curvilinear
flight path, we obtain the following equations of motion
from Newton’s second law.
• Along the flight path:
𝑑𝑉
• 𝑊𝑠𝑖𝑛𝜃 − 𝐷 = 𝑚 𝑑𝑡
• Perpendicular to flight path:
𝑉2
• 𝐿 − 𝑊𝑐𝑜𝑠𝜃 = −𝑚 𝑅
• R  local radius of curvature of the flight path
Hypersonic Flight Paths: Velocity-Altitude Maps
• For most entry conditions, θ is small, hence we assume sin 𝜃 ≈ 0 𝑎𝑛𝑑 𝑐𝑜𝑠𝜃 ≈ 1.
• Noting that m = W/g
𝑊 𝑑𝑉
• −𝐷 =
𝑔 𝑑𝑡
𝑊 𝑉2
• 𝐿−𝑊 = −
𝑔 𝑅
• The drag can be expressed in terms of the drag coefficient 𝐶𝐷 𝑎𝑠 𝐷 = 12𝜌𝑉 2 𝑆𝐶𝐷
• Where ρ  freestream density, S  reference area
𝑊 𝑑𝑉
•  −12𝜌𝑉 2 𝑆𝐶𝐷 =
𝑔 𝑑𝑡
• Rearranging,
1 𝑑𝑉 𝑊 −1 𝜌𝑉 2
• − =
𝑔 𝑑𝑡 𝐶𝐷 𝑆 2
𝑊
• is defined as the ballistic parameter; it clearly influences the flight path of the entry vehicle via the solution of
𝐶𝐷 𝑆
above equation
𝑊
• For a purely ballistic reentry (no lift), is the only parameter governing the flight path for a given entry angle.
𝐶𝐷 𝑆
Hypersonic Flight Paths: Velocity-Altitude Maps
• Expressing the lift in terms of the lift coefficient 𝐶𝐿 𝑎𝑠 𝐿 = 12𝜌𝑉 2 𝑆𝐶𝐿 , we obtain
𝑊 𝑉2
• 𝐿−𝑊 = −𝑔 𝑅
𝑊 𝑉2
• 1
2
𝜌𝑉 2 𝑆𝐶𝐿 −𝑊 = −
𝑔 𝑅

• Rearranging,
1 𝑉2 𝑊 −1 𝜌𝑉 2
• 1− 𝑔 𝑅 = 𝐶𝐿 𝑆 2
𝑊
• is the lift parameter; it clearly influences the flight path of a lifting-entry vehicle via the solution of above Eq.
𝐶𝐿 𝑆
𝑊 𝑊
• Lift parameter and Ballistic parameter equations illustrate the importance of 𝐶 𝑆 and 𝐶 𝑆 in determining the flight
𝐷 𝐿
path through the atmosphere of a vehicle returning from space
• Such flight paths are frequently plotted on a graph of altitude vs velocity—a velocity-altitude map
Hypersonic Flight Paths: Velocity-Altitude Maps
• Such flight paths are frequently plotted on a graph of altitude vs velocity—a velocity-altitude map, an
example of which is shown in Fig
• Here, two classes of flight paths are shown:
𝑊
1) lifting entry, governed mainly by 𝐶 𝑆,
𝐿
𝑊
2) ballistic entry, governed mainly by .
𝐶𝐷 𝑆

• The vehicle enters the atmosphere at


either satellite velocity (such as from orbit)
or at escape velocity (such as a return from a
lunar mission).
• As it flies deeper into the atmosphere, it
slows as a result of aerodynamic drag,
giving rise to flight pathsshown in Fig
Hypersonic Flight Paths: Velocity-Altitude Maps
𝑊 𝑊
• Note that vehicles with larger values of 𝐶 and/or 𝐶 penetrate deeper into the atmosphere before slowing.
𝐿𝑆 𝐷𝑆
𝑊
• The lifting-entry curve for 𝐶 = 100 lb/ft2 pertains approximately to the space shuttle;
𝐿𝑆
𝑊
• The curve initiated at escape velocity with 𝐶 𝑆 = 100 lb/ft2 pertains approximately to the Apollo entry
𝐷
capsule.
• Velocity-altitude maps are convenient diagrams to illustrate various aerothermodynamic regimes of
supersonic flight.
Basic Hypersonic Shock Relations
• Anytime a supersonic flow is turned into itself (such as flowing over a wedge, cone, or
compression corner), a shock wave is created.
• Also, if a sufficiently high backpressure is created downstream of a supersonic flow, a standing
shock wave can be established.
• Such shock waves are extremely thin regions (on the order of 10-5 cm in air) across which large
changes in density, pressure, velocity, etc. occur.
• These changes take place in a continuous fashion within the shock wave itself, where viscosity and
thermal conduction are important mechanisms.
• However, because the wave is usually so thin, to the macroscopic observer the changes appear to
take place discontinuously.
• Therefore, in conventional supersonic aerodynamics shock waves are usually treated as
mathematical and physical discontinuities.
• As the Mach number is increased to hypersonic speeds, no dramatic qualitative difference occurs.
• However, some interesting approximate and simplified forms of the shock relations are obtained in
the limit of high Mach number
Basic Hypersonic Shock Relations
Basic Hypersonic Shock Relations
• Standard plot for wave angle vs. deflection
angle with Mach number as a parameter.
(θ – β – M diagram)
• From this figure, note that, in the
hypersonic limit, when θ is small, β is
also small.
Basic Hypersonic Shock Relations

Applying high Mach number limit


Hypersonic Shock Relations in Terms of the
Hypersonic Similarity Parameter
• In the study of hypersonic flow over slender bodies, the product 𝑀1 𝜃 is an important governing parameter,
where, as before, 𝑀1 is the freestream Mach number and 𝜃 is the flow deflection angle.
• 𝑀1 𝜃 is a similarity parameter for such flows.
• Denoting 𝑀1 𝜃 by K, we state

• Return to the exact θ – β – M relation given by

• As expressed, this is an explicit relation for 𝜃 = 𝜃 𝛽 . Obtaining the exact inverse relation, 𝛽 = 𝛽 𝜃 ,
from above equation is tedious.
Hypersonic Shock Relations in Terms of the Hypersonic
Similarity Parameter
• However, in the combined limit of hypersonic flow and small angles, an approximate explicit relation for 𝛽 =
𝛽 𝜃 can be obtained.
• For small angles above equation reduces to

• Rewritten as,

• Rearranging,
Hypersonic Shock Relations in Terms of the Hypersonic
Similarity Parameter
Hypersonic Expansion-Wave Relations
• M1, p1  Mach number and pressure upstream of the Prandtl-Meyer expansion wave
• M2, p2  Mach number and pressure downstream of the wave
• From compressible flow, the relation between θ, M1 and M2 is given by
𝜃 = 𝜈 𝑀2 − 𝜈 𝑀1
• Where ν  Prandtl-Meyer function
𝛾+1 𝛾+1 2
𝜈 𝑀 = tan −1 (𝑀 − 1 ) − tan−1 𝑀2 − 1
𝛾−1 𝛾−1

You might also like