19.01.2015 Views

11. Hypersonic Aerodynamics - the AOE home page - Virginia Tech

11. Hypersonic Aerodynamics - the AOE home page - Virginia Tech

11. Hypersonic Aerodynamics - the AOE home page - Virginia Tech

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>11.</strong>1 Introduction<br />

<strong>11.</strong> <strong>Hypersonic</strong> <strong>Aerodynamics</strong><br />

<strong>Hypersonic</strong> vehicles are commonplace. There are more of <strong>the</strong>m than <strong>the</strong> true supersonic<br />

aircraft discussed in <strong>the</strong> last chapter. Applications include missiles, launch vehicles and entry<br />

bodies. Never<strong>the</strong>less, <strong>the</strong>re is no exact definition defining <strong>the</strong> start of <strong>the</strong> hypersonic flow<br />

regime. Possibilities are:<br />

a) Mach numbers at which supersonic linear <strong>the</strong>ory fails<br />

b) Where γ is no longer constant, and we must consider temperature effects on<br />

fluid properties.<br />

c) Mach numbers from 3 - 5, where Mach 3 might be required for blunt bodies<br />

causing large disturbances to <strong>the</strong> flow, and Mach 5 might be <strong>the</strong> starting point<br />

for more highly streamlined bodies.<br />

In this section we will provide a brief outline of <strong>the</strong> key distinguishing concepts. The book by<br />

Anderson 1 provides a starting point for fur<strong>the</strong>r study.<br />

Essentially <strong>the</strong>re are five key points to be made:<br />

1. Temperature and aerodynamic heating become critically important<br />

2. In many cases pressure can be estimated fairly easily<br />

3. Blunt shapes are commonplace<br />

4. Control and stability issues lead to different shapes<br />

5. Engine-airframe Integration is critical<br />

<strong>11.</strong>2 Importance of Temperature/Aerodynamic Heating<br />

This is easily illustrated considering <strong>the</strong> relation between stagnation temperature and static<br />

temperature:<br />

Or <strong>the</strong> wall temperature<br />

T 0<br />

T<br />

T adiabatic<br />

wall<br />

γ −1 2<br />

=1 + M<br />

2 ∞<br />

⎛<br />

= ⎜ 1+ r γ −1<br />

2 M ∞ 2<br />

⎝<br />

(11-1)<br />

⎞<br />

⎟ T<br />

⎠ e (11-2)<br />

where r is <strong>the</strong> recovery factor. With <strong>the</strong>se relations, <strong>the</strong> limit for aluminum structure is around<br />

Mach 2, which was <strong>the</strong> Concorde’s cruise Mach number. The SR-71 is made of titanium, and<br />

temperature limits <strong>the</strong> speed to slightly over Mach 3. Note that some sources indicate that this<br />

limit is actually <strong>the</strong> temperature limit on <strong>the</strong> wiring in <strong>the</strong> airplane, and that is <strong>the</strong> condition<br />

that limited <strong>the</strong> speed.<br />

Thus hypersonic aerodynamic configuration design means that you must always deal with<br />

heating. In general, at sustained high speeds surfaces must be cooled, and since heating is a<br />

critical concern, this means that viscous effects are crucial immediately. Also, unlike normal<br />

5/8/09 11-1


11-2 Configuration <strong>Aerodynamics</strong><br />

airplane aerodynamics, hypersonic vehicles fly at very high altitudes, and <strong>the</strong> Reynolds<br />

number may be so low that <strong>the</strong> flow is laminar. Thus laminar flows are often of interest. This<br />

is important because <strong>the</strong> heat transfer is much lower when <strong>the</strong> flow is laminar. In fact, being<br />

able to estimate <strong>the</strong> transition location with certainty is a critical requirement in hypersonic<br />

vehicle design, and is <strong>the</strong> subject of current research. 2<br />

The X-15, shown in Figure 11-1, is <strong>the</strong> only real manned hypersonic airplane flown to<br />

date. It was rocket powered, and started flight by being dropped from a B-52, so it was purely<br />

a research airplane. The first flight was by Scott Crossfield in June of 1959. The X-15 reached<br />

314,750 feet in July of 1962. An improved version reached a Mach number of 6.7 and an<br />

altitude of 102,100 feet in October of 1967 with Pete Knight at <strong>the</strong> controls. The X-15<br />

program flew 199 flights, with <strong>the</strong> last one being in October of 1968. * Milt Thompson’s book 3<br />

describes <strong>the</strong> X-15 program, including <strong>the</strong> crackling sounds in <strong>the</strong> airframe as it heated up!<br />

Figure 11-1. X-15 (NASA Dryden photo archives photo)<br />

In case <strong>the</strong> issue of aerodynamic heating seems academic, consider <strong>the</strong> M = 6.7 flight of<br />

<strong>the</strong> X-15. It turned out to be <strong>the</strong> last flight of that plane. A dummy ramjet installation was<br />

* The only undergraduate from <strong>the</strong> <strong>Virginia</strong> <strong>Tech</strong> Aerospace Engineering program to earn astronaut status was<br />

Jack McKay, who flew <strong>the</strong> X-15 as a NASA test pilot, and reached 295,000 feet in September of 1965.<br />

5/8/09


W.H. Mason <strong>Hypersonic</strong> <strong>Aerodynamics</strong> 11-3<br />

tested by placing <strong>the</strong> ramjet mockup below <strong>the</strong> airplane. The shocks from <strong>the</strong> front of <strong>the</strong><br />

ramjet inlet spike impinged on <strong>the</strong> pylon supporting <strong>the</strong> scramjet. The shock interference<br />

heating was so severe that <strong>the</strong> shocks acted as a blowtorch, cutting through <strong>the</strong> structure, and<br />

effectively slicing off <strong>the</strong> scramjet. The internal damage to <strong>the</strong> airplane from <strong>the</strong> heating led to<br />

<strong>the</strong> scrapping of this high-speed version of <strong>the</strong> airplane. Figure 11-2 below shows <strong>the</strong> scramjet<br />

hanging below <strong>the</strong> airplane.<br />

Figure 11-2. X-15 with dummy ramjet (Picture from <strong>the</strong> NASA Dryden Photo Web Site)<br />

<strong>11.</strong>3 Surface pressure estimation<br />

In many cases surface pressures are relatively easy to estimate at hypersonic speeds. At<br />

supersonic speed we have a local relation for two-dimensional flows relating surface slope<br />

and pressure:<br />

C p =<br />

2θ<br />

M 2 −1<br />

(11-3)<br />

However, this relation is not particularly useful in most cases on actual aircraft<br />

configurations. In comparison, hypersonic rules are useful. The most famous relation is based<br />

on <strong>the</strong> concepts of Newton. Although Newton was wrong for low-speed flow, his idea does<br />

apply at hypersonic speeds. The idea is that <strong>the</strong> oncoming flow can be though of as a stream<br />

of particles, which lose all <strong>the</strong>ir momentum normal to a surface when <strong>the</strong>y “hit” <strong>the</strong> surface.<br />

This leads to <strong>the</strong> relation:<br />

C p = 2sin 2 θ (11-4)<br />

where θ is <strong>the</strong> angle between <strong>the</strong> flow vector and <strong>the</strong> surface. Thus you only need to know <strong>the</strong><br />

geometry of <strong>the</strong> body locally to estimate <strong>the</strong> local surface pressure. Also, particles impact only<br />

<strong>the</strong> portion of <strong>the</strong> body facing <strong>the</strong> flow, as shown in Figure 11-3. The rest of <strong>the</strong> body is in a<br />

“shadow”, and <strong>the</strong> Cp is zero. See Anderson 1 for <strong>the</strong> derivation of this and o<strong>the</strong>r pressureslope<br />

rules.<br />

Two key observations come from <strong>the</strong> Newtonian rule. First, <strong>the</strong> Mach number does not<br />

appear! Second, <strong>the</strong> pressure is related to <strong>the</strong> square of <strong>the</strong> inclination angle and not linearly as<br />

it is in <strong>the</strong> supersonic formula. This illustrates how <strong>the</strong> situation in hypersonic flow is<br />

significantly different than <strong>the</strong> linearized flow models at lower speed.<br />

5/8/09


11-4 Configuration <strong>Aerodynamics</strong><br />

oncoming<br />

flow<br />

Shadow<br />

Region,<br />

Cp = 0.0<br />

Figure 11-3. Shadow sketch, showing region where Cp is zero<br />

This figure also shows how <strong>the</strong> situation changes from low speed to high speed. Look at <strong>the</strong><br />

“ultimate” lift plot in Figure 11-4. Here <strong>the</strong> lower surface pressure is taken equal to <strong>the</strong><br />

stagnation value, and <strong>the</strong> upper surface is taken equal to <strong>the</strong> vacuum value. We can see that at<br />

low speeds <strong>the</strong> lift is generated on <strong>the</strong> upper surface, while at high speed <strong>the</strong> lift is almost<br />

completely generated on <strong>the</strong> lower surface.<br />

CL<br />

ultimate<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

"Ultimate" CL<br />

Note <strong>the</strong> switch from upper surface<br />

to lower surface dominated, and <strong>the</strong><br />

relatively low values from Mach 0.9<br />

to Mach 1.5.<br />

6.0<br />

Total Lift<br />

4.0<br />

Upper<br />

Surface<br />

2.0<br />

Lower<br />

Surface<br />

0.0<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Mach<br />

Figure 11-4. The “Ultimate CL” plot (showing <strong>the</strong> dominance of <strong>the</strong> lower<br />

surface at hypersonic speeds)<br />

The Newtonian flow model can be refined to improve agreement with data. This form is<br />

known as <strong>the</strong> Modified Newtonian flow formula,<br />

where <strong>the</strong> stagnation Cp is a function of Mach and γ,<br />

C p = C pmax sin 2 θ (11-5)<br />

C pmax = p 0 2<br />

− p ∞<br />

1 2<br />

ρ ∞ V ∞<br />

2 (11-6)<br />

and P 02 is <strong>the</strong> stagnation or total pressure behind a normal shock. This expression gets both <strong>the</strong><br />

Mach number and ratio of specific heats back into <strong>the</strong> problem. The classical Newtonian<br />

5/8/09


W.H. Mason <strong>Hypersonic</strong> <strong>Aerodynamics</strong> 11-5<br />

<strong>the</strong>ory is actually <strong>the</strong> limit as M → ∞, and γ → 1. As briefly noted above, <strong>the</strong>re are lots of<br />

o<strong>the</strong>r local rules, and Anderson’s book 1 should be consulted for a complete discussion.<br />

<strong>11.</strong>4 Round LE shapes to reduce heat transfer<br />

Since round shapes have lots of drag, why do we see such highly rounded shapes The answer<br />

is <strong>the</strong> heat transfer. The maximum value of <strong>the</strong> heat transfer is related to <strong>the</strong> leading edge<br />

radius as shown in this relation:<br />

q ˙ max ≈<br />

laminar<br />

1<br />

R LE<br />

(11-7)<br />

where R LE is <strong>the</strong> leading edge radius at <strong>the</strong> stagnation point.<br />

Thus, we need a nose or leading edge radius large enough for <strong>the</strong> tip not to melt. H. Julian<br />

Allen and A.J. Eggers, Jr., at NASA Ames did <strong>the</strong> analysis that showed that blunt bodies were<br />

required to survive entry from orbit. 4 Thus, <strong>the</strong> solution of hypersonic flows about <strong>the</strong> nose of<br />

an entering body became an important analysis problem. In fact, this was <strong>the</strong> first great<br />

challenge problem for CFD (this was in <strong>the</strong> 1960s, and <strong>the</strong> computational solution of <strong>the</strong><br />

flowfield wasn’t called CFD until <strong>the</strong> early 1970s). It is known as “<strong>the</strong> Blunt Body problem”.<br />

It was particularly difficult because <strong>the</strong> flow is subsonic behind a normal, or nearly normal,<br />

shock wave. The flow <strong>the</strong>n accelerates quickly to supersonic/hypersonic speed. This is <strong>the</strong><br />

reverse problem of transonic flow, where <strong>the</strong> freestream is supersonic/hypersonic ra<strong>the</strong>r than<br />

subsonic. In particular, we want to know <strong>the</strong> shock standoff distance, <strong>the</strong> shock shape, and <strong>the</strong><br />

flow properties at <strong>the</strong> nose, where <strong>the</strong> aerodynamic heating is largest. The shape of <strong>the</strong> shock<br />

determines <strong>the</strong> distribution of flow properties such as entropy, which vary as <strong>the</strong> shock shape<br />

changes.<br />

The problem was solved by computing <strong>the</strong> unsteady flowfield, which is always<br />

ma<strong>the</strong>matically hyperbolic. If <strong>the</strong> solution is actually steady, <strong>the</strong> computation will converge to<br />

<strong>the</strong> steady state result, while overcoming <strong>the</strong> difficulty of <strong>the</strong> mixed elliptic-hyperbolic<br />

equation type that describes <strong>the</strong> steady state problem. The successful method invented to<br />

obtain a practical blunt body calculation method is generally attributed to Moretti. 5<br />

<strong>11.</strong>5. Aerodynamic control<br />

We described <strong>the</strong> benefits of blunt shapes above. Many hypersonic vehicles display ano<strong>the</strong>r<br />

unusual geometric feature. It turns out that thick bases are often used on hypersonic vehicles,<br />

and in this section we illustrate why <strong>the</strong>y are desirable. Our example relates to <strong>the</strong> thick<br />

trailing edge on <strong>the</strong> vertical tail of <strong>the</strong> X-15.<br />

The example of <strong>the</strong> difference in <strong>the</strong> flow characteristics between supersonic and<br />

hypersonic speed can be illustrated with <strong>the</strong> hypersonic directional stability problem. To start,<br />

we consider <strong>the</strong> yawing moment, which is<br />

where <strong>the</strong> standard definition of C N is:<br />

N VT = q VT S VT l VT C YVT (11-8)<br />

C n =<br />

N<br />

q ref S ref b ref<br />

(11-9)<br />

5/8/09


11-6 Configuration <strong>Aerodynamics</strong><br />

We can write <strong>the</strong> yawing moment due to <strong>the</strong> vertical tail as:<br />

C nVT =<br />

l VT S VT<br />

b ref S<br />

ref<br />

<br />

vertical tail<br />

volume<br />

coefficient, V VT<br />

q VT<br />

⋅ C YVT (11-10)<br />

q<br />

ref<br />

ratio of<br />

dynamic pressures<br />

assume ≈ 1<br />

The nomenclature associated with <strong>the</strong> problem and <strong>the</strong> equations is illustrated in Figure 11-5.<br />

⋅<br />

β<br />

V<br />

∞<br />

N<br />

l VT<br />

Y VT<br />

US<br />

LS<br />

θ<br />

Figure 11-5. Sketch defining <strong>the</strong> hypersonic directional stability problem<br />

Now, for a high-speed flow, we will assume that <strong>the</strong> vertical tail is a two-dimensional<br />

surface with a constant pressure on each side, so that C YVT = C PLS − C PUS . We will consider<br />

two cases, one supersonic, <strong>the</strong> o<strong>the</strong>r hypersonic. We compare <strong>the</strong> results for directional<br />

stability at high Mach number using <strong>the</strong> two-dimensional rule for linearized supersonic flow<br />

and Newtonian <strong>the</strong>ory for hypersonic flow, as shown below.<br />

θ<br />

5/8/09


W.H. Mason <strong>Hypersonic</strong> <strong>Aerodynamics</strong> 11-7<br />

Linear <strong>the</strong>ory<br />

2θ<br />

C p =<br />

M 2 −1<br />

Newtonian <strong>the</strong>ory<br />

C p = 2sin 2 θ (11-11)<br />

Case 1: Linearized supersonic <strong>the</strong>ory<br />

( )<br />

( )<br />

C YVT = ΔC p = 2 θ + β<br />

M 2 −1 − 2 θ − β<br />

M 2 −1 =<br />

showing that <strong>the</strong> θ’s cancel. We use this expression to get C nβ :<br />

C nβ<br />

VT = V VT<br />

∂C YVT<br />

∂β<br />

= V VT<br />

4<br />

M 2 −1<br />

4β<br />

M 2 −1<br />

This expression shows that C nβ is positive, but vanishes for hypersonic Mach numbers.<br />

(11-12)<br />

(11-13)<br />

Case 2: <strong>Hypersonic</strong> flow <strong>the</strong>ory, Newtonian <strong>the</strong>ory<br />

This time <strong>the</strong> expression for <strong>the</strong> side force is:<br />

Using trig functions:<br />

Which reduces to:<br />

and<br />

C Yβ<br />

VT<br />

C YVT = ΔC p = 2sin 2 ( θ + β) − 2sin 2 ( θ − β) (11-14)<br />

⎡<br />

= 2 sin2 θ cos 2 β + 2cosθ sinβ sinθ cos β + cos 2 θ sin 2 β<br />

⎤<br />

⎢<br />

⎥<br />

⎢<br />

⎣ −sin 2 θ cos 2 β + 2 cosθ sin βsinθ cos β − cos 2 θ sin 2 ⎥<br />

β⎦<br />

C YβVT<br />

= 8cosθ sin β<br />

sinθ cosβ <br />

≈1 ≈β ≈θ ≈1<br />

≅ 8θβ<br />

(11-15)<br />

(11-16)<br />

C nβ<br />

VT = 8V VTθ (11-17)<br />

If θ is zero, so is C nβ ! But opening up <strong>the</strong> wedge angle rapidly increases C nβ . In addition,<br />

<strong>the</strong>re is no Mach number dependence. The Case 2 results were verified experimentally, and<br />

<strong>the</strong> wedge vertical tail concept literally saved <strong>the</strong> X-15 program. 6 This effect is also <strong>the</strong> reason<br />

for flared “skirts” seen on some launch vehicles.<br />

The application of this effect is clearly shown in <strong>the</strong> first picture of <strong>the</strong> X-15 above and <strong>the</strong><br />

front view of <strong>the</strong> X-15 shown in Figure 11-6 below.<br />

5/8/09


11-8 Configuration <strong>Aerodynamics</strong><br />

Figure 11-6. Picture of <strong>the</strong> X-15 showing <strong>the</strong> wedge used for <strong>the</strong> vertical tail “airfoil”<br />

section. (from <strong>the</strong> US Air Force Museum Web Site)<br />

<strong>11.</strong>6. Engine-airframe Integration<br />

Any air-breathing hypersonic vehicle will have a highly integrated engine and airframe. It will<br />

be hard to tell exactly what counts as <strong>the</strong> vehicle, and what counts as <strong>the</strong> propulsion system.<br />

This is exemplified in <strong>the</strong> concepts developed for recent aerospace planes. In <strong>the</strong>se concepts<br />

<strong>the</strong> propulsion is at least in part provided by a scramjet engine, which obtains thrust with a<br />

combustion chamber in which <strong>the</strong> flow is supersonic. Figure 11-7, from a presentation by<br />

Walt Engelund of NASA Langley, shows this concept. The entire forebody of <strong>the</strong> vehicle<br />

underside is used as an external inlet to provide flow at just <strong>the</strong> right conditions to <strong>the</strong> engine.<br />

The entire underside afterbody is <strong>the</strong> exhaust nozzle.<br />

The first US vehicle to use a scramjet engine was <strong>the</strong> Hyper-X, now known as <strong>the</strong> X-43.<br />

The initial flight attempt was made in June of 2001, but failed before having a chance to<br />

demonstrate <strong>the</strong> operation of <strong>the</strong> scramjet. After a long investigation and a revised procedure,<br />

it flew successfully on March 27, 2004. 7 The X-43A is shown in <strong>the</strong> Figures 11-8 and 11-9.<br />

The November-December 2001 issue of <strong>the</strong> Journal of Spacecraft and Rockets had a<br />

special section devoted to <strong>the</strong> Hyper-X.<br />

5/8/09


W.H. Mason <strong>Hypersonic</strong> <strong>Aerodynamics</strong> 11-9<br />

Figure 11-7. Example of engine-airframe integration for a hypersonic aircraft<br />

(from Walt Engelund, NASA Langley, May 2001).<br />

Figure 11-8. Artist conception of <strong>the</strong> X-43<br />

5/8/09


11-10 Configuration <strong>Aerodynamics</strong><br />

Figure 11-9. The X-43 at NASA Dryden in flight preparation early in 2001.<br />

<strong>11.</strong>7 Airbreathing <strong>Hypersonic</strong> Vehicle Design<br />

The successful design of an airbreathing hypersonic vehicle is an excellent example of <strong>the</strong><br />

need for multidisciplinary optimization methods. This is a very hard problem. The recent<br />

success of <strong>the</strong> scramjet test flight means that we can expect to see designs proposed using this<br />

technology. Key surveys are available based on numerous studies conducted at NASA<br />

Langley. 8,9<br />

Finally, we note that one concept not discussed so far deserves mention. Waveriders<br />

potentially provide efficient hypersonic flight. These designs can be thought of as placing a<br />

shape in <strong>the</strong> streamline of body generated flowfield so that <strong>the</strong> bottom surface is resting on <strong>the</strong><br />

pressure field generated by <strong>the</strong> shock wave of this flowfield. The arrangement is designed to<br />

obtain lift with very low drag. 10<br />

<strong>11.</strong>10 Exercises<br />

1. Derive expressions for <strong>the</strong> lift curve slope of a flat plate and a wedge using linearized<br />

supersonic <strong>the</strong>ory and Newtonian <strong>the</strong>ory. Comment on <strong>the</strong> differences.<br />

<strong>11.</strong>9 References<br />

1 John Anderson, <strong>Hypersonic</strong> and High Temperature Gas Dynamics, McGraw-Hill, New<br />

York, 1989. (reissued as a second edition by <strong>the</strong> AIAA)<br />

2 John J. Bertin and Russell M. Cummings, “Fifty years of hypersonics: where we’ve been,<br />

where we’re going”, Progress in Aerospace Sciences, Vol. 39, pp. 511-536, 2003 (Note:<br />

ignore <strong>the</strong> error in <strong>the</strong> first line of <strong>the</strong> abstract, that has <strong>the</strong> wrong date and pilot for <strong>the</strong> Mach<br />

6.7 flight)<br />

5/8/09


W.H. Mason <strong>Hypersonic</strong> <strong>Aerodynamics</strong> 11-11<br />

3 Milton O. Thompson, At <strong>the</strong> Edge of Space: The X-15 Flight Program, Smithsonian<br />

Institution Press, Washington, 1992.<br />

4 H. Julian Allen and A.J. Eggers, Jr., “A Study of <strong>the</strong> Motion and Aerodynamic Heating of<br />

Ballistic Missiles Entering <strong>the</strong> Earth’s Atmosphere at High Supersonic Speeds,” NACA<br />

Report 1381, 1958.<br />

5 Gino Moretti and Mike Abbett, “A Time Dependent Computational Method for Blunt Body<br />

Flows,” AIAA Journal, Vol. 4, No. 12, 1966, pp. 2136-2141.<br />

6 Charles McLellan, “A Method for Increasing <strong>the</strong> Effectiveness of Stabilizing Surfaces at<br />

High Supersonic Mach Numbers,” NACA RM L54F21, Aug. 1954, Declassified: 1958.<br />

7 Michael A. Dornheim, “A Breath of Fast Air,” Aviation Week & Space <strong>Tech</strong>nology, April 5,<br />

2004.<br />

8 Mary K. O’Neill and Mark J. Lewis, “Design Tradeoffs on Engine-Integrated <strong>Hypersonic</strong><br />

Vehicles,” AIAA Paper 92-1205, Feb. 1992.<br />

9<br />

Mary K. Lockwood, Dennis H. Petley, John G. Martin and James L. Hunt, “Airbreathing<br />

hypersonic vehicle design and analysis methods and interactions,” Progress in Aerospace<br />

Sciences, Vol. 35, pp. 1-32, 1999.<br />

10 Charles E. Cockrell, Jr., Lawrence D. Huebner and Dennis B. Finley, “Aerodynamic<br />

Characteristics of Two Waverider-Derived <strong>Hypersonic</strong> Cruise Configurations,” NASA TP<br />

3559, July 1996.<br />

5/8/09

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!