Вы находитесь на странице: 1из 17

Physics of High Speed Flows (I)

Contents
1-

Physics of High Speed Flows (I)....................................................................................................... 2

1.1 -

Atmospheric (RE-)Entry physics .................................................................................................. 2

1.2 -

The ballistic coefficient and the trajectory and vehicle design..................................................... 4

1.3 -

A typical example ........................................................................................................................ 12

1.4 -

Major aerothermodynamic phenomena of re-entry vehicles .................................................... 13

1.5 -

References ................................................................................................................................... 17

1-

PHYSICS OF HIGH SPEED FLOWS (I)

1.1 -

Atmospheric (RE-)Entry physics

An important step forward for Space Exploration activities and for a more accurate
knowledge of the Earth, and Universe is to develop the capability to send vehicles into space
which select, collect and finally return samples from other celestial bodies to Earth where to
perform their analysis. There are therefore, space missions that require the spacecraft (S/C) to
go from space through the atmosphere to the surface (of Earth or another planet). This is
currently called Entry and this process may be:

re-entry, if the S/C is coming back to Earth, or

entry, if the S/C is arriving and landing on another planet.


In what follows we are going to refer to the re-entry process only.
Remark: There are also other types of flights/situations in which similar physical processes
occur and this are the cases of the descent of ballistic missiles and/or of the sounding
(suborbital) rockets.

In each case, there are the following main aspects to look at (because the mission of the S/C
have to be accomplished and this is a must):

Deceleration. The S/C payload (including crew) and structure limit the maximum
deceleration (measured in gs) allowed. This deceleration must be: a) low enough to
prevent damages or injuries to occur (for instance, the trained human payload can
endure without special equipments loads of about +/- 12-g for short periods and the
structure may be designed consequently) and b) high enough to brake the S/C to (re)enter rather than to skip off the atmosphere. There are some missions designed so that
the deceleration must be realized without any active devices (like rocket motors and/or
parachutes).

Heating. The deceleration is associated with an intense heating caused by the heat
exchange between the S/C and the air molecules. The heat flux transferred to the S/C is
so important that the heating of the S/C may damaged or even destroyed. The entry
trajectory and de thermal protection system of the S/C must be designed so that to
prevent the effects of the heating.

Accuracy of landing (or impact). The S/C will be expecting to arrive at an imposed
location on the planet, with a given accuracy. This influences both the design of the S/C
and entry trajectory.

The deceleration, heating and accuracy requirements define a permissible (re-)entry corridor,
see Figure 2.1. If the S/C does not enter such a corridor, then it will experience either too much
heating and burn or too little drag and skip off the atmosphere. The size of the corridor
depends on three factors, which are: deceleration, heating and (landing) accuracy.

Figure 2.1 A typical reentry corridor for a S/C.


The space vehicle configuration and the entry trajectory are the result of a trade-off process,
see Figure 2.2, which is based on the mission requirements for the S/C. Mission requirements
affect directly the vehicle design.

Figure 2.2 The mission requirements determine the re-entry design process.
Remark. A high speed Earth re-entry vehicle has the following characteristics:

entry velocity higher or equal to 11.7 km/s (compared to 7.5 km/s for the Space
Shuttle),
very high heat fluxes (more than 10 MW/m2) and
heat loads (in the range of 500 MJ/m2), where the radiative part is important.
The design of such a S/C relies on a good understanding of the aerodynamic loads encountered
during the atmospheric part of the re-entry. The shape of such a capsule is the result of a tradeoff design among hypersonic aerodynamic loads (aeroheating), subsonic drag and subsonic
stability.

1.2 -

The ballistic coefficient and the trajectory and vehicle design

The forces acting on the S/C are the weight and the aerodynamic drag and lift, see Figure
2.3. During the initial phase of the entry, one may assume that the aerodynamic drag is
dominant. The flight-path angle, , is the angle between the local horizontal and the velocity
vector. The entry trajectory is characterized by the initial velocity and the entry flight-path
angle, while the S/C shape and size and the thermal protection systems are characterizing the
vehicle design.
1

If the drag force is given by an equation like = 2 2 , where the drag coefficient CD

depends on the vehicle shape and flying conditions and S is the (cross) reference area, then the
acceleration on the trajectory is given by:
1 2

= 2 .

2.1)

The ballistic coefficient is defined as the denominator of the equation (2.1),


=

2.2)

Figure 2.3 The forces acting on the spacecraft.


From equation (2.1) one sees that the magnitude of the deceleration from drag is
inversely related to BC. The BC quantifies an S/C mass, drag coefficient and (cross-sectional)
area and predicts how drag will affect it. The Figure 2.4 shows an intuitive relation between the
shape of the vehicle and its BC. As a result, the deceleration depends on BC and an S/C with
small ballistic coefficient (a blunt body) will decelerate much stronger than one with a high BC
(a streamlined body).

Figure 2.4 The S/C shape and its ballistic coefficient.


The trajectory design is based on controlling the re-entry initial conditions, i.e. the
velocity and the flight path angle with which the S/C enters the atmosphere. For most missions,
the re-entry initial conditions are set by the mission orbit and are difficult or even impossible to
change significantly (without using propulsion).
The vehicle design is based on: a) choosing the S/C shape and size, see Figure 2.5, so
that to have a convenient BC and after then, b) providing a thermal protection system to deal

with entry heating. The shape and size determine the magnitude of the aerodynamic forces and
thus the ballistic coefficient.

Figure 2.5 Various S/C shapes and sizes and the corresponding BCs.
The deceleration has a maximum value, that depends on the initial velocity and flight path
angle, Figure 2.6. For a given re-entry flight-path angle, the higher the initial velocity, the
greater the maximum acceleration. Further, for a given velocity, the higher the re-entry flightpath angle, the greater the maximum acceleration. The acceleration and its maximum value
determine the time that an entering S/C takes to get down to dense layers of the atmosphere.
Using the flight dynamics equations one can find the S/Cs maximum deceleration and
the altitude at which it occurs. These are given by:

sin

, where

2.3)

= 0.000139 1 (atmospheric scale height) and = 2.7182 (base of the natural logarithm).
and

= ln 0sin , where

2.4)

0 = 1.225 /3 (air density at the sea level).


The maximum deceleration depends on the initial re-entry velocity and flight-path angle.
However, the altitude of the maximum deceleration depends only on the flight-path angle, as
can be noticed from (2.4).

Figure 2.6 Deceleration profiles for various re-entry velocities and flight-path angles.
Because of the extremely high re-entry velocities, the travel through the upper
atmosphere produces significant aero-thermodynamic effects and these are related to the
occurrence of the very strong shock waves, see Figure 2.7. The hot air downstream of the shock
wave transfers some of its heat to the S/C by convection. The convection is known to be the
primary means of heat transfer to a S/C entering the Earths atmosphere at speeds under about
15000 m/s. Above this speed, the air molecules are so hot that they transfer more of their
energy to the S/C by radiation. The heat transfer process is quantified through the heating rate
(heat energy per unit area per unit tome), , measured in watts per square meter.
For streamlined bodies the shock wave attaches to the tip and thus a significant
amount of heat transfers to the body in this region, see Figure 2.8. These may cause a strong

localized heating of the S/Cs sharp tip and thus a very high temperature may be reached.
Furthermore, the air flow near the surface does not restrain the heat transfer and as a result,
the overall heating rate is high, as it is shown in Figure 2.9 (streamlined vehicles have a large
BC).

Figure 2.7 Attached and detached shock waves,


for high (streamlined) and low (blunt) BC vehicles.
If the vehicle is blunt, the shock wave detaches and curves in front of the vehicle, leaving a
boundary of air between the shock wave and the S/Cs surface. This spreads the heat over a
larger volume of air. The air flow near the the surface tends to slow down the convective heat
transfer. Thus, the heating rate for blunt vehicles is relatively low.

Figure 2.8 The impact of the shape on the heating process.


Figure 2.9 shows how different BC affect the maximum heating rate. It is a matter of evidence
that the heating rate is more severe and occurs much lower in the atmosphere for high BC
vehicles (i.e., for streamlined S/C with sharp tip). This is due to the shape of the shock wave, as
it was explained previously.

Figure 2.9 The dependence of the heating rate on the BC.


The heating rate is a function of the S/Cs velocity and nose radius ( ) and also to
the density of the atmosphere. An empirical formula for the heating rate shows that:

= 1.83 104 3
.

The equation (2.1) confirms that the smaller the nose radius, the higher the heating rate.
Figure 2.10Figure 2.10 shows a typical variation of the heating rate with the altitude
and also the dependency of the peak heating rate value on the re-entry velocity and flight-path
angle.

Figure 2.10 Variation of the heating rate with altitude,


at different re-entry velocities (left) and flight-path angles (right).

2.5)

Using some algebraic manipulations one may find the velocity and the altitude where the
maximum heating rate occurs. Thus, this velocity is given by

and the corresponding altitude is

= 0.846 ,

= ln 30sin , where

2.6)

2.7)

= 0.000139 1 (atmospheric scale height) and


0 = 1.225 /3 (density at the sea level).

From Figure 2.10 and the previous equations results that:


the maximum heating rate increases as the re-entry velocity increases,
the velocity for the maximum heating rate is about 85% of the re-entry velocity,
the steeper the re-entry angle, the higher the maximum heating rate.
A steep re-entry angle causes a very high heating rate but for a short time, so the overall
effect on the vehicle may be small. A low re-entry angle leads to much lower heating rates, but
acting for longer times. The total heat load, Q, is the amount of thermal energy per unit surface
(J/m2) the S/C receives and is obtained by integrating the heating rate over the entire re-entry
time.

Figure 2.11 The total heat load for various re-entry velocities.

Q varies therefore with the velocity but it does not vary with the re-entry flight-path
angle (the heat results from the mechanical energy dissipation process during re-entry). Figure
2.11 shows an example from which one may observe that the higher the re-entry velocity, the
higher the total heat load is. Although the peak heating rate depends on the flight-path angle,
the total heat load is constant for a given re-entry velocity.

Table 2.1. Ballistic Coefficient (BC) trade-offs for the re-entry

In what is concerning the accuracy, it must be noticed that the aerodynamic drag
and lift forces perturb the descending trajectory from the path it would follow under gravity
alone. The modeling of these forces (including the density of high atmosphere and the
dimensionless coefficients) can be done only with approximations. To reduce these
atmospheric effects on the accuracy, the trajectory must be so that the S/C spends the least
time in the atmosphere. This leads to high re-entry velocities and steep re-entry flight-path
angles and clearly, this choice increases the severity of the deceleration and heating rate.
The effects of the BC on the re-entry process are summarized in Table 2.1.
The important heat accumulation that occurs during the re-entry is counteracted by
thermal protection systems. These are specially materials and design technique and include:

Heat sinks (the heat is spread out and stored in the S/C, using an extra (quantity of)
material to absorb the heat, thus keeping the peak temperature lower);
Ablation (the re-entry vehicle is sheltered with a material having a very high latent heat
of fusion, which absorbs through melting and/or vaporization a large amount of heat
energy and disappears in atmosphere);
Radiative cooling (the re-entry vehicle is sheltered with a material that has a high
emissivity and thus it reaches the thermal equilibrium at a relatively low temperature by
emitting back to the atmosphere almost as much heat energy as it absorbs).
The heat sinks are simple but heavy and they reduce the payload mass significantly. The
ablation process can be used only once, because the S/C must be completely renewed for each
mission. The radiative cooling requires a surface coating with high emissivity and melting point

and also an inner efficient insulator to protect the structure, but is the only solution used for
reusable vehicles (the Space Shuttle, for instance).
The BC, the shape and size of the S/C, the parameters of the re-entry trajectory
and the thermal protection system represent the factors that determine the physical processes
that take place during the re-entry. Correspondingly, they also determine the mathematical
models that may/must be used for the accurate and efficient quantitative prediction of the flow
and heat transfer by means of numerical computation.

Remark on lifting re-entry. The use of the lift force offers more flexibility in the re-entry
trajectory design and also opens the possibility of aerobraking, which is an exciting application
mainly for the entry in the atmospheres of another planets. However, this does not change the
facts presented previously. In other words, the braking in the atmosphere, the heating and the
need for thermal protection system still remain the essential factors of the re-entry process.

1.3 -

A typical example

We consider in what follows a typical flight scenario for a sample return mission. The reentry trajectory is high speed one, with a steep descent characterized by an initial flight path
angle of about -12.5 deg. The flight scenario is shown in Figure 2.12. Such a trajectory example:

allows predicting the flow field around the proposed capsule because of it
provides freestream conditions for CFD computations,
defines the aerothermodynamic environment the capsule has to withstand
during descent. For instance, the evolutions of heat fluxes (convective and
radiative part) vs. the velocity are presented.
In the preliminary design the convective and radiative heat fluxes have been estimated by
using analytical engineering correlations such as Scott relationship for convective heat flux and
Tauber-Sutton one for the radiative one. These estimations are important for designing of the
S/C heat shield, because the aeroheating environment dictates, in fact, the type and size of the
thermal protection system to use.

Figure 2.12 Altitude vs. velocity and heating flux evolution for a typical sample return mission.
The peak heat rate generally determines the range of possible thermal protection material
while the integrated heat load determines the thickness and hence the mass of the heat shield.
Furthermore, the strong flowfield radiation and heatshield ablation determine the return
vehicles aerothermodynamic performances.
Due to the very high temperatures reached in the shock layer caused by the
strong bow shock in front of the capsule, the gas dissociates and ionized and also emits
radiated energy. This energy travels through the flowfield and interacts with the gas itself. At
the entry velocity predicted for this sample return mission this contribution, that is generally
very low with respect to other energies in the flowfield, cannot be neglected because it can
cause an additional source of heating load at the wall to be taken into account. The flowfield
and the radiative heat field are coupled. Mathematically, this means that in the Navier-Stokes
equations a source term is added to the energy equation to take into account the energy
radiated from and towards the vehicle walls.

1.4 -

Major aerothermodynamic phenomena of re-entry vehicles

Aerothermodynamics, or hypersonic aerodynamics in the frame of this work, is based on


mathematical models that take into account the specific physical processes. When modeling
these physical processes it is necessary to specify the class of the vehicles on which the focus is.
Although there are numerous common aerothermodynamic features to all the hypersonic
vehicles there are also very different key technology demands for different vehicle classes. In
the frame of the present work we discern three classes of hypersonic vehicles:

winged return vehicles (RV), like the Space Shuttle, BURAN and HERMES,
ascent and re-entry vehicles (ARV) and
aero-assisted orbital transfer vehicles (AOTV).

Each of these three classes has special aerothermodynamic features, which lead to different
research, mathematical modeling and development needs. The features are presented in Table
2.2.

From the classification results that the following physical effects play a major role:
viscosity effects, notably laminar-turbulent transition and turbulence,
thermodynamic effects, especially plasma (ionization, radiation),
heat loads and the cooling concept.
A review of the modeling and simulation means of the potential critically aerothermodynamic
phenomena is presented in Table 2.3.

Table 2.3 shows that there are improvements to be done for the modeling and simulation of all
the enumerated physical phenomena. The simulation accuracy demand for each phenomena
depends on the sensitivity of the vehicle or its components on the respective phenomena.

Table 2.2. Comparative consideration of the aerothermodynamic features of the major classes
of re-entry vehicles
RV
ARV
AOTV
Mach number range

28-0

0/7-28

20-35

Configuration

blunt

Opposite demands at blunt


ascent and re-entry
(sharp/blunt)

Flight time through short


atmosphere

short

short

Angle of attack

large

small/large

head on

Drag

large

small/large

large

Lift to drag ratio

small but larger than small


the others

Thermal
problems

heating loads

small

loads

loads

Flowfield

pressure
dominated

field viscous
effects pressure
dominated / pressure dominated
field dominated

Rarefaction effects

initially strong

weak / initially strong

strong

Thermodynamic
effects

strong

medium/strong

strong

Critical components

control
surfaces, nozzle/base,
landing syst.
surfaces

Special problems

large Mach number Propulsion


span
integration,
demands

control control devices


Plasma effects
opposite

field

Table 2.3. Review of the modeling and simulation means for aerothermodynamic phenomena
Phenomena
Modelling and simulation Vehicle class
means
Transition laminar-turbulent

poor

RV, ARV

Attached turbulent flow

good

RV, ARV, AOTV

Laminar strong interaction

good

RV, ARV, AOTV

Turbulent strong interaction

fair

RV, ARV

Laminar separation

good

RV, AOTV

Turbulent separation

fair

RV, ARV

Hypersonic viscous interaction good


(low density effects)

RV, ARV, AOTV

Equilibrium real-gas effects

good

RV, ARV, AOTV

real-gas good

RV, ARV, AOTV

Non-equilibrium
effects

Turbulent heat transfer

poor

RV, ARV

Surface radiation cooling

Good (in USA)

RV, ARV, AOTV

Plasma effects

fair

AOTV

1.5 -

References

1. Sellers, J. and all, Understanding Space: An Introduction to Astronautics, McGraw-Hill, 2000.


2. Aerothermodynamic Field Past a Reentry Capsule for sample Return Missions, by A. Viviani,
G. Pezzella, C. Golia, 28TH INTERNATIONAL CONGRESS OF THE AERONAUTICAL SCIENCES,
ICAS 2012.
3. Hypersonic Aerodynamics, by E.H. Hirschel, Deutsche Aerospace AG, Space Course 1993, TU
Munich, Oct. 11-22, 1993.
4. Stagnation-Point Radiative Heating Relations for Earth and Mars Entries, by M.E. Tauber, K.
Suttont, Journal of Spacecraft and Rockets, VOL. 28, NO. 1 (pag.40-42).
5. Thermal Design of Aeroassisted Orbital Transfer Vehicles, by C.D. Scott, R.C. Ried,
C.P. Li and S.M. Derry, "An AOTV Aeroheating and Thermal Protection Study," H. F.
Nelson (ed.), Thermal, Vol. 96 of Progress in Astronautics and Aeronautics, AIAA,
New York, 1985, pp. 198-229.

Вам также может понравиться