• Nie Znaleziono Wyników

Supersonic flow over low aspect-ration wings

N/A
N/A
Protected

Academic year: 2021

Share "Supersonic flow over low aspect-ration wings"

Copied!
91
0
0

Pełen tekst

(1)

r

C » " C V V N \ U , M . hf

Kiuyvèrweo 1 - 2823 Hi,

SUPERSONIC FLOW OVER LOW ASPECT-RATIO WINGS

TECHNISCh'f HOGESCHOOL DELFT

LUCHTVAART- EH ilUliVTEVAARTTECHNIEK BIBLIOTHEEK

Kluyverweg 1 - DELFT

by

\l

\jit-V^n

I.e. Richards, M.A.(Oxon.)

Submitted for the degree of Doctor of Philosophy of the Cranfield Institute of Technology

Cranfield Institute of Technology College of Aeronautics

Aerodynamics Division Cranfield

(2)

Sumnary

List of Figures

Notation

Introduction

Acknowledgments

Part 1. The Design Condition for a Caret Wing i

1.1 Introduction 1

1.2 Flow over a caret wing j,

1.3 Design condition 1

m Detachment condition 2

1.5 Strong and weak shock solutions 3

1.6 The experimental facility 3

1.7 Tests and results

n

Part 2. The Anomalous Leeside Pressure Rise 6

2.1 Introduction 6

2.2 Test procedure and results 6

2.3 Further tests 8

Part 3. The Supersonic Delta Wing 10

3.1.1 Introduction 10

3.1.2 Flow regimes on delta wings 10

3.1.2a Thin wings 10

3.1.2b Effect of thickness 13

3.1.2c Effect of leading-edge radius m

3.1.2d Effect of Reynolds Niunber 15

3.2.1 Thin shock-layer theory 16

3.2.2 Extension to the prediction of leeside flows 25

3.3 Tunnel investigation of a particxilar delta wing 26

3.3.1 Aims of the investigation 26

3.3.2 The models 27

3.3.3 Test procedure 28

(3)

(List of C o n t e n t s - c o n t i n u e d ) 3.3.3a P r e s s u r e P l o t t i n g 3.3.3b S c h l i e r e n p h o t o g r a p h s 3.3.3c V a p o u r s c r e e n t e c h n i q u e 3.3.3d O i l f l o w t e s t s 3.3.4 R e s u l t s 3 . 3 . If a F l o w o v e r t h e c o m p r e s s i o n s u r f a c e 3.3.fb F l o w o v e r the e x p a n s i o n s u r f a c e 3.3.5 C o n c l u s ions 3.4 T h e i m p o r t a n c e o f t h e n u m e r i c a l s c h e m e in t h i n s h o c k - l a y e r t h e o r y 3.5 F u r t h e r e x p e r i m e n t s t o v a l i d a t e t h e a m e n d e d t h e o r y 3.5.1 A i m and d e s c r i p t i o n o f t e s t s 3.5.2 R e s u l t s 3.5.3 C o n c l u s i o n s 3.6 F i n a l c o n c l u s i o n s and s u g g e s t i o n s f o r f u r t h e r w o r k 28 28 28 29 29 29 30 33 33 35 35 36 37 37 R e f e r e n c e s F i g u r e s

(4)

consolidate and augment existing knowledge of high-speed flows over delta wings. Particular emphasis is placed on the investigation of flows which did not conform to the 'accepted' pattern.

In Part 1 the design conditions of a caret wing are discussed and references in the literature to 'strong oblique-shock solutions' on such wings are shown to be misleading. The results of a tunnel investigation of a caret wing designed to support a 'strong-shock' are presented.

In Part 2 the unexpected pressure rises reported on the lee surfaces of various delta wings (references 1 and 2) are shown to be, at least in part, the result of interference from the model support and base-mounted instrumentation. 'Correct' leeside pressure distributions are presented for one of the models used in reference 2, for angles of attack up to 50 degrees.

In Part 3 the different flow regimes on delta wings are discussed together with the methods of defining the boundaries between them. The conjecture that thin shock-layer theory can be used to predict the onset of leading-edge sepciration is carefully investigated by means of tunnel tests on a particular wing with triangular cross-section. The data are presented as upper and lower surface pressure distributions and as oil flow, vapour screen and schlieren photographs, at angles of attack up to 50 degrees. A later version of thin

shock-layer theory, which employs a physically more acceptable method to solve the governing equation, is also described. The predictions for shock wave angle, streamline pattern and pressxu<e distributions are compared with experiments for a wing with flat compression surface. Suggestions for further work are made.

(5)

LIST OF FIGURES

Figure

The testing of a delta wing Frontispiece Flow regimes on a typical cau?et wing

Coordinate system

Caret wing with a dual design-condition at M - 2*45 Caret wing designed to support a strong, oblique shock Schlieren photographs of 'strong-shock' caret wing Pressure distributions on 'strong-shock' caret wing Details of flat-cone model

Model support system in C.I.T. tunnel

Flat cone with original sting and pressure tubes Effect of removing flexible tubes from base of model Model with angled sting

Leeside pressure distributions at x/i =0*55

Leeside pressure distributions at V i - 0*55 (after Szodruch and Squire)

IH Attempts to promote leeside pressure rise 15 Model with 20°, 20% chord flap

16 Model with wedge in wake

17 Construction of effective Mach Number and incidence 18 Flow^regimes on thin delta wings

19 Effect of thickness on shock detachment 20 Flow regimes on wings of finite thickness

21 The experimental boundary between regimes of attached and separated leeside flows (after Stanbrook and Squire) 22 Variation of inverse density ratio (e) with incidence 23 Predictions of Shanbhag's theory for thin wings

(after Squire)

24 Predictions of the 'half-modified' theory for the model under investigation

25 Dimensions of the delta wing models 26 The models

27 The working-section and pressure-scanning switch 28 Vapour screen picture, showing major features

(continued) 1 2 3 4 5 6 7 8 9 10 U 12 13 \

(6)

37 Combined compression-sxirface pressure distributions

38 Schlieren photographs

-5^ f a i

50°

39 Variation of shock wave angle with incidence

40 Oil flow photographs

41-48 Expansion-surface pressure distributions

49 Variation of pressure with chordwise position

50 Vapour screen photographs

51 Variation of vortex height with incidence

52 The propagation of disturbances in thin shock-layer

theory

53 Oil-dot flow on flat compression surface

54 Movement of compression-surface attachment lines

55 Vapour screen photographs

(7)

N O T A T I O N

Cp Non dimensional pressure coefficient

Cp(vac) Pressure coefficient 'in vacuo' = ~ /yM^

M Mach Number

P Pressure

P.^ Total pressure

£ Velocity vector

Re£ Reynolds Number based on model length

T^ Total temperature

x, y, z Natural coordinates of figure 2

u, V, w Velocity components

a Angle of incidence relative to leading-edge plane

Op Semi-vertex angle of wedge, angle of incidence of lower

ridgeline of caret wing

6

(M2

- l)i

Y Ratio of specific heats

X Sweepback angle in planform (caret wing, a^ = 0)

A Sweepback angle in leading-edge plane

p Density

a

Local body inclination

w Angle of 'undercut' (caret wing)

Subscripts

b value at wing surface

s value insiediately behind shock

oo value in undisturbed stream

(8)

The work reported in this thesis was carried out under the Ministry of Defence, Procurement Executive contract entitled 'The Flow over Low Aspect-Ratio Wings at Supersonic and Hypersonic Speeds'. The Author duly records his thanks.

Throughout the tenure of his contract the Author was supervised by Professor J.L. Stollery, who always found time to take an active part in the research. For his advice and guidance, which have proved literally invalucüale, the Author is deeply grateful.

Regular meetings with the Contract Monitor, Mr. P.L. Roe, and Dr. L.C. Squire, have provided the theoretical 'back-up' necessary for any research programme, and the form in which the work appears is in no small measure due to their assistance. The same may be said of Professor B.A. Woods, whose timely currival from New Zealand enabled the last sections of this thesis to be written.

Especially when he first arrives, a student is deeply dependent on the help of the technical staff and in this, the Author feels that he has been unusually fortunate. For steering him around some pitfalls and patiently watching him drop into others, he would like to thank the Division's workshop staff, especially Mr. D.F. Sibley and Mr. M.F. Goodridge but above all Mr. S.W. Clarke who, amongst many other things, missed his lunch probably more times than he cares to remember, in order to run the supersonic tunnel.

The Author would like to thank Mrs. J. Holloway for her careful and accurate typing in this and previous papers.

This thesis is dedicated to Mr. J.R. Busing who died shortly

before it was completed and in whom much of the Division's experimental expertise resided. His help and friendship will be sadly missed.

(9)

I N T R O D U C T I O N

The delta wing with conical thickness distribution represents a highly idealised approximation to a real supersonic aircraft wing. Nevertheless many of the features of the flow on these simple wings are mizn?ored in the full scale flows. For this reason, the delta wing has come to play much the same role in high speed research as has the two dimensional aerofoil in the low-speed case.

The College of Aeronautics at Cranfield has two major high-speed facilities, a continuous supersonic tunnel with a working section measuring 230 x 230 mm and an intermittent hypersonic helium tunnel with a core diameter of 100 mm. Together, the two tunnels cover the Mach number range 1>6 < M^ ( 30.

In the case of the delta wing research described here, it was deemed desirable to consolidate existing knowledge of the subject, by attempting to answer specific questions, rather than to add slavishly to the available data. In order that the results of the work should be pertinent to current and next-generation aircraft, it was also decided to restrict the experimental work to supersonic Mach numbers.

Three main areas of supersonic flow on delta wings were thought to require systematic investigation:

1. Strong and weak shock solutions for a caret wing

Confusing references to 'strong oblique-shocks' are sometimes found in the literature when discussing the design condition of a caret wing. A more careful investigation usually reveals that the choice of words has been inappropriate and no analogy with the strong shock solution in wedge flow is intended. Altho\igh never observed in an unbounded, two dimensional flow, steady, strong

oblique-shocks have been observed in external compression jet intakes. It was surmised that a suitably designed caret wing might also

support such a shock, its position being stabilised by the presence of the leading edges.

(10)

are usually close to the vacuum limit and eu?e often neglected in the calculation of aerodynamic loads. Experimental work by Hillier^ and by Szodruch and Squire^ however, showed that under certain

conditions leeside pressures could rise above that of the freestream at moderate incidences. They were unable to suggest, except in a very general way, the reason for this pressure rise.

There was (and is) no mathematical model for separated leeside flows capable of giving an accurate prediction of pressure

distributions or vortex position. In spite of the foregoing remarks regarding the low values of leeside pressure, knowledge of the vortex position and strength is important in the positioning of leeside stabilisers and control surfaces and in the prediction of heat transfer rate.

3. Onset of leading-edge separation

A rather diffuse and restricted empirical boundary separates regimes of attached and separated leeside flows. Squire^ has attempted to use a simple theory to predict changes in the

compression svirface flow which may acccntipany leading-edge separation on the lee surface. The theory needed experimental verification, especially f ^ wings of finite thickness.

(11)

1

P A R T 1

The Design Condition for a Caret Wing

1.1 Introduction

The caret wing is the simplest of the family of wings known as 'wave riders' (so called because their shape is designed to support a given, usually simple, shock configuration). It has an inverted 'V' cross-section and a delta planform. For particular combinations of Mach number and incidence the wing will support a two-dimensional flow field with a plane shock attached to the leading edges. The design condition may be readily calculated from the oblique shock relations and therefore most of the reseaz*ch effort has been directed towards the 'off-design' behaviour. Some caret wings exhibit two design incidences for a given Mach number and this sometimes causes confusion in the literature.

1.2 The flow over a caret wing

Figure 1 shows the different flow regimes on a typical caret wing. ST is the design curve and PR the shock detachment

boundary. Starting at zero incidence and fixed Mach number of say M Q , we see that at first the Shockwave is attached to the leading edges but bowed outwards (region B ) . As we cross the design curve the shock becomes plane and the flow behind it, two dimensional. Further increase of incidence brings us through region A, where a

'branched' shock is necessary to satisfy the equations of motion, back to the design curve. Above this Incidence the shock again bows outwards (region C) and finally detaches from the leading edges

(region D ) . For the zero-thickness case shown, the curves ST and PR can be obtained from the oblique shock relations as shown below (see figure 2 for notation):

1.3 Design condition

A caret wing 'on-design' produces a two dinenslonal flow field; the generating 'wedge' is the wing ridgeline and the Shockwave is required to lie in the plane of the leading edges. It is therefore

(12)

only

necBBBary

to solve the oblique shock relation, for the wedge

angle Op ^^^^ the shock angle 9,

- « • o

M 2

Sin^ e-1

tan

Op -

Cot e ;; ^

, .

(Y+l)Mi/2 - ( M 2 Sin2 e-1)

together with the condition that 6 - Op = Const = w, to produce

the design curve. The solution is conveniently fo\md by graphical

means; for example from reference 4.

1.4 Detachment condition

By resolving the incident flow into a plane normal to the

leading-edge, the oblique shock relationships may be applied to the

'equivalent wedge' in order to produce the exact shock detachment

bo\indcu?y. With reference to figure 2, we obtain the following

relationships for the normal components of Mach nuinber and incidence:

Mn = M (1 - Cos^o Sin^ A ) ' (1.2)

c,.„

_ Sin (g - Ü») Cos A

, .

Sin One 1 ' ' 1 tl.3;

(1 - Cos^w Sin2 A ) * (1 - Cos^o Sin^ A ) *

For a given a we may calculate anc And hence from reference 4,

the corresponding Mach number for shock detachment ( M - ) . Equation

(1.2) then yields the free stream Mach number for shock detachment.

The results obtained from these 'exact Inviscid' calculations

are In good agreement with experiment.

The leeside flow on caret wings has received little specific

attention. Volume is usually added to the wing in such a way as to

produce streamwise upper surfaces when the wing is 'on-design'. At

other incidences the leeside flows are similar to those on

(13)

3

1.5 Strong and weak shock solutions

In cases where a dual design condition exists, the two solutions are sometimes referred to as the 'weak' and 'strong' solutions. This may lead lis to believe (and indeed we are sometimes told explicitly, for instance in reference 5) that the solutions are analogous to the

'weak' and 'strong' solutions for the shock waves on a wedge,

predicted by the oblique shock relations. Gonor^ justifies his use of the terms by saying that the downstream coo^onent of Mach number normal to the shock is subsonic in the 'strong' solution. He then goes on to say that this solution is never observed in the flow over a wedge, which is not true.

For the caret wing shown in figure 3, we see that for a Mach number of 2*45 there aro two design incidences, viz. Oj* ~ ^^ ^^^

Op - 20^. Reference to the graph in the same figure shows that both these solutions correspond to the 'weak' branch of the 6 vs. or curve, i.e. in both cases the flow is supersonic behind the shock.

Although a steady, strong oblique-shock never occurs on a simple wedge, Neale and Lambe^ at the NGTE report the existence of a strong shock in external compression intakes where the flow is substantially two dimensional. In order to see if a caret wing could support a single strong shock, a steeply «nhedralled wing was constructed and tested in the Cranfield supersonic tunnel. Details of the model and its design curve are shown in figxire 4. The

planform of this wing is the same as that of the 'conventional' caret wing of figure 3. The design incidence at M„ s 2*45 is also the

same {op - 20°) but the two design and detachment curves are otherwise quite different.

1.6 The experimental facility

The supersonic tunnel used at Cranfield is of the closed circuit, continuous variety. Interchangeable liners provide a Mach number range of 1*6 to 2«8 in discrete steps. The stagnation pressure can be varied from zero to one atmosphere by controlling the quantity of air admitted to the tunnel after starting. The 230 mm x 230 mm working section hoxises a crescent model moxmting

(14)

which allows a continuous adjustment of incidence in the range -5° t a i 25°. The sting length is chosen to place some part of the model at the centre of rotation.

Throughout the tests described in this thesis the Mach number was kept constant at a nominal 2*5. The actual test conditions were as follows: H^ = 2-45 Pt = 3*5 X lO*» Nm"2 5 psla Re^ s 4 X io5 T^ = Ambient Adiabatic wall

Model pressxires were at first measured by means of vertical merctiry manometers but these were found to be too insensitive. For the caret wing tests, ethylene glycol was used in the manometers. Although suiteible in its low specific gravity and high boiling point, the high viscosity seriously reduced the rate at which readings

couJ.d be taken.

A change to a pressure scanning switch and data logger

subsequently improved both the speed and accuracy of the measuz>ements. A conventional single-pass schlieren system was eiq>loyed.

1.7 Tests and resvilts

The results of the model tests are shown in figures 5 and 6 as a series of schlieren photographs and pressure distributions covering the incidence range 10° < op < 30°. Although not easy to interpret explicitly, the schlieren photographs clearly show a complex shock 8tz>ucture; probably of the 'branched' variety as observed by Crabtree and Treadgold^ and typical of a caret wing below design incidence (region 'A' in figure 1 ) . The pressure distributions tell a similar story, though the grid is too coarse to fix accurately the shock position.

A 'spot' result obtained by Roe^ using thin shock-layer theory gives the following predictions for ox> = 27*28°, which are

(15)

5

compared with the linearly interpolated experimental results.

T.S.L.T. Expt. Incidence «r 27*28 27*28 Angle between inboard shock and ridgeline 36*78 deg. 32 deg. P/P. (ridgeline) 6*86 5*7 P/P. (outboard region) 1*54 2*5

The agreement is rather striking in view of the low Mach number and unmodelled viscous effects, and leaves little doubt that the shock structure is similar to that predicted by thin shock-layer theory.

The conclusion drawn is that a simple caret wing cannot support a single strong oblique shock. Because the flow behind such a shock is subsonic, downstream conditions become of great importance and it is thought to be the careful control of these conditions which allowed Neale and Lambe to obtain the steady, strong solution in their intake tests.

(16)

P A R T 2

The Anomalous Leeside Pressure Rise

2.1 Introduction

Hillier^, and Szodruch and Squire^ have tested families of wings with delta planforms and have observed unusual trends in the development of leeside flows with increasing Incidence. As the incidence increased from zero, the leeside pz^essure at first fall as ejected but then, at about 12 degrees, began to rise and by 14 degrees had reached the free stream static value. Subsequent increases in incidence produced further small rises in pressure. Complementary schlieren photographs showed this pressure rise to be accompanied by a movement o'f the vortex away from the lee surface and the appearance of a shock from the apex of the wing. The same qualitative behaviour was shown when the models were tested in the 3 ft X 4 ft supersonic ttinnel at the Royal Aircraft Establishment, Bedford.

At the suggestion of L.C. Squire, one of the models was

brought to Cranfield to investigate the effect of support interference on the leeside flow. Details of the model are shown in figure 7.

2.2 Test procedure and results

The mounting lug on the model proved incompatible with the model support system in the 9 in. K| 9 in. tunnel and was therefore removed and replaced by a sting as shown in figure 8. The pressure plotting tubes were unchanged, and those on the flat upper surface were connected by 3 nn O.D. PVC tubing to the tunnel pressure measuring system. The result of this initial test showed suction Increasing normally xxp to an Incidence of about 20 degrees. At a - 23*5 degrees the flow became highly unsteady with large pressxire fluctuations on the upper surface. At a = 24 degrees, a steady shock appeared over the upper surface springing from the apex of the wing and the pressvire rose above the free stream static level. A series of schlieren photographs in figure 9 illustrates this sequence.

(17)

7

The length of the sting was then increased by 70 mm; this produced no detectable change in the flow pattern, and indicated that it was not the proximity of the incidence crescent which affected the flow.

Finally the flexible tubing was removed from the back of the model and the free ends of the hypodermic tube bent slightly inwards. A comparative schlieren photograph (figure 10) taken at a s 2(1 degrees shows a significemt change in the flow with no evidence of the apex shock observed earlier. A second sting was then constructed to allow the incidence to be increased to 50 degrees; a single, small diameter flexible tube was used to monitor the pressure on the leeside

centreline. Schlieren photographs (figure 11) show no abrupt change in the flow pattern; the apex shock finally appearing at about

a - 35 degz>ee8. Further thin flexible tubes were then cautiously joined to the hypodermic tubes. Schlieren photographs indicated that the flow had not been significantly disturbed. With the aid of these smaller tubes some pressure distributions were measxired; the results for 55% chord are shown in figure 12 for the Incidence range 0 $ a < 50 degrees. As can be seen, the pressure falls to approximately 0.7 Cp(vac) at a = 35 degrees and thereafter rises gradually to reach the free stream static value at a = 50 degrees. There were no significant differences in the results for 80% chord. Figure J.3 shows a specimen result from Squire and Szodznich's investigation for comparison. The particular test was conducted at M : 3*5 and thez*efore the numerical value of the pressure coefficients cannot be directly compared with those in the previous figure. The divergence in the results for a > 12 degrees is clearly visible however.

An inspection of the schlieren photographs shows that the leeside vortices 'bend' away from the wing surface at some point. This 'bend', which Szodruch^° associates with vortex burst, lies close to the trailing edge at moderate incidences and begins to move upstream at about a s 20 degrees. The measured pressure distributions though indicate that the pressure rise occurs simultaneously at the two chordwise stations. More chordwise

(18)

pressure measurements would be necessary to clarify the connection between the state of the vortex and the leeside pressure.

2.3 Further tests

The explanation for the leeside pressure rise observed in the earlier tests at the R.A.E. and Cambridge would appear to be a large scale flow separation resulting from a severe adverse pressure gradient over some part of the sup[>oz*t. To produce artificially such a

pz>essure gradient, three rectangular flaps were constructed (figure 14) and tested successively on the model. Flap A was set at zero

incidence (relative to the lee s\irface) and extended backwards a distance equal to 20% root chord. Flaps B and C were 20 degree wedges of 10% and 20% root chord respectively. They extended forward from the trailing edge of the model.

The model was re-tested in the range 20° i a ( '50°. Flap A was found to have little effect, the apex shock and pz>es8ure rise occ\u?ring much as before at a = 35 degrees. Flaps B and C produced local regions of separated flow in the compression comer (figure 15) but otherwise had little effect on the leeside flow. In particular there was no evidence of the separated region extending upstream with increasing incidence.

At this stage it was decided to restore the model to its

original configuration, i.e. as in figure 9, to check that the initial observationscould be reproduced. At the maximum Incidence

obtainable with the straight sting (a s 25 degrees) thez*e was no evidence of leeside flow separation. This was irritating rather than disturbing. In the original tests the sudden pressure rise had occurred very close to the limit of incidence adjustment and the critical incidence was known to be very sensitive to small changes in the model.

An examination of the model revêked that at some time during the tests the tip had become rather blunted and this was therefore built up and re-sharpened; a suggested mechanism for the observed pressure rise had been asymmetrical vortex shedding which is known to depend strongly on the tip geometry. In this case however, the

(19)

9

sharpening was not found to have any effect. Finally, a large wedge was placed in the wake of the model just behind and above the trailing edge (figure 14D). (Szodruch^° had found this effective in promoting the leeside pressure rise.) A schlieren photograph (figure 16) taken at a s 25 degrees shows no evidence that the flow over the wing has been substantially altered.

At this stage it was concluded that the anomalous leeside flows observed by earlier workers were the result, at least in part, of the model support and instrumentation. The exact mechanism(s) still had to be pinpointed but vortex burst seemed an unlikely contender.

Since these tests were completed, undergraduates at Cambridge^^ have investigated the effect of boundary layer transition on the

critical incidence for leeside pressure rise. Their results indicated that the critical incidence could be raised by several degrees by

means of boundary layer trips on either the windward or leeward surfaces. It should not surprise us that the leeside flow is affected by the tripping of the compression surface boundary layer since when the shock wave is detached turbulence may be convected around the leading edge. What is so sxirprising is that the state of the boundary layer should so pz>ofouiklly affect the development of the leeside flow. The usual effect of boundary layer transition

(see section 3.1.2a) is merely to displace slightly the position of the secondary*vortex. It is not known what Interpretation should be put on this finding; taken jointly the eiqperimental results show no systematic dependence of critical incidence upon Reynolds Number.

(20)

P A R T 3

The Supersonic Delta Wing

3.1.1 Introduction

In this section, two pieces of e3q>erimental work are described together with two versions of thin shock-layer theory from

substantially independent authors. Had the later version of the theory been available when the experimental programme was being planned, its content would have been markedly altered. In order to preserve the correct sequence of events therefoz>e, the work presented here is arranged in chronological order.

( • 3.1.2 Flow regimes on delta wings

a) Thin wings i

The supersonic flow over a delta wing is complicated and depends on sweepback, thickness, leading-edge radius, incidence, Mach number and Reynolds Number. In order to reduce the number of variables and to develop a 'feel' for the type of flows

encountered, we will initially restrict our discussion to thin wings with sharp leading edges. This permits a further

simplification into two significant parameters, viz. the component of Mach number normal to the leading edge (M^) and the flow

inclination in this plane ( O Q ) . These are related to sweep. Mach number and incidence by:

and Mn - M,(l - Cos^ a Sin^ A ) ^ (3.2)

(see figure 17). With the aid of this redefinition of variables, the data for thin wings may be represented on a single graph (see figure 18), and three flow regimes identified,corresponding to:

(21)

11

a) detached shock on compression surface / separated flow on expansion surface b) detached shock on compression surface /

attached flow on expansion surface c) attached shock on compression surface /

attached flow on expansion surface

The shock detachment boundary separating regions B and C is obtained by applying the two dimensional oblique shock relations in a plane normal to the leading edges (as in section 1.4). The detachment condition is 'exact' and may be obtained from either a complicated analytical expression or from shock charts (e.g. reference 4 ) . It will be noted that shock detachment can occur either as a result of the leading edges of the wing becoming subsonic (i.e. M^^ < 1) or because the equivalent wedge angle (an) exceeds the detachment angle.

Accompanyiiig a detached shock there may be either attached or separated flow au?ound the leading edge (regions B and A in

figure 18 respectively). In region B the flow remains attached around the leading edge, turning by means of a Prandtl Meyer expansion. The streeunlines on the leeside are directed towards the centreline and there will in general be a shock to z>eturn the flow to the free-stream direction. This shock may separate the boundary layer and result in the formation of an inboard vortex. In region A, the leeside flow is similar to the 'low-speed' case. The boundary layer separates at the leading edge and rolls up to form a vortex lying above the wing. There will in general be an attachment line iiiboard, with a region of streamwise flow between this and the wing centreline. With increasing

incidence this region will diminish in extent until the attachment line reaches the centreline.

As in the low-speed case there may be further separation outboard of the primary attachment line, with an attendant counter-rotating,secondary vortex.

In region C, leading-edge separation is not possible and the leeside flow is qualitatively similar to that in region B.

(22)

The boundary between regions A and B, shown cross-hatched in figure 16, was proposed by Stanbrook and Squlz^e^^ from their compilation of delta wing data published originally in 1952 (see figure 21). To the left of the boundary (region A) no instances of attached leeside flows had been recorded and to the right of the boundary (region B) no instances of sepetrated flows. Within the boundary region both separated and attached flow had been

observed (Including mixed flows on a single wing). Although the data used by Stanbrook and Squire are for 'thin' wings, this does not imply that the wings had zero thickness. This finite thickness, of which no account was taken in reference 12, could at least partly explain the diffuse nature of the boundary; a plot in (an, Mn)

(where an is measured relative to the leading-edge plane] cannot be e]q>ected to correlate data for a range of wing thicknesses.

Since 1952 the Squire-Stanbrook boundary has been somewhat extended in the light of further experimental work but there is currently no theoretical model for leeside flows capable of predicting its position.

In 1975 Squire^ suggested that the onset of leading edge separation would be accompanied by marked changes in the compression surface flow. He used thin shock-layer theory to predict the

conditions under which such changes may occur. This will be discussed in^detail in Section 3.2.

The exact mechanism by which leading edge separation occurs is also open to conjecture. According to Squire^ the sequence of events in crossing from region B to region A (say by reducing the Mach niamber) is as follows:- In region B in figure 18, the flow expands around the leading edge in a Prandtl Meyer wave and there will in general be a region of constant pressure on the leeside, terminated by a shock. As the Nach number falls the flow is unable to expand completely and leading edge separation occurs, though the vortex remains close to the wing surface cuid the inboard flow is still'terminated by a shock. With further decreases in Mach number the expansion fan fades out leaving a fully developed vortex flow. .

(23)

In Rein's^^ model, the region of uniform pressure inboard of the leading-edge is explained by the presence of a leeside separation bubble which forms as soon as the shock detaches from the lower surface. This bubble is analogous to that observed on a wedge in transonic flow and is assumed to be of constant width. It therefore interacts with the inboard shock near the apex, resulting in a vortex being shed over the inner part of the wing. As the Mach number falls, the bubble increases in width and the vortex in extent. Eventually the region of attached flow inboard of the bubble vanishes and the leading edge vortex covers the whole wing.

Both these models have corroborative experimental evidence though Rein's results could also be explained by a 'transition' argument in which the leuninar boundary layer near the apex is separated by the inboard compression, whereas the turbulent layer farther downstream is not.

b) Effect of thickness

As long as the leading edge remains sharp the two-dimensional shock detachment conditions can be easily modified to take account of finite wing thickness. Assuming a and On are defined as before relative to the plane of the leading edges and that, in the normal plane, the wing has thickness 6^ ëübove and 6i below the leading edge^plane (see figure 19) then the results for a family of wings will collapse in a plot of (a^^ t ój^) vs. Mjj for

(a^ t S^) i (6„ t 6i)/2. The equality sign represents the

condition that the 'equivalent wedge' is symmetric with respect to the direction of the normal component of the free stream.

For normal incidences below this, the conventionally 'leeward' surface controls the shock detachment and a family of curves result. The effective incidence of the leeside plane is given by (6^ ~ On) and it is this value of a (effective)which must be used in the equation (e.g. reference 14) for shock detachment. Some examples are shown in figure 20.

(24)

For incidences above On + 6i = (^n •*" * A ) / 2 . Squire^ finds that the ((a^j •»- 6^), MjJ representation satisfactorily correlates the available data for leading-edge separation. The apparently weak dependence of the leeside flow on leeside shape is interesting and supports the view that changes in the compression surface flow are responsible for leeside separation. It should be noted that even with a detached shock wave, the upper surface cannot influence the lower surface flow (beyond the boundary layer) so long as there exists a sonic line between the shock euid the leading-edge region of the wing, and so long as the changes to the leeside shape are not so gross as to change significantly the position of this sonic line.

c) Effect of leading-edge radius

In the early days of delta wing research, much attention was focussed on obtaining a 'classical' subsonic type of flow and the wings therefore had a conventional aerofoil section with a well-rounded leading edge and a sharp trailing edge. Later, it became apparent that leading-edge separation was inevitably going to occur somewhere in the flight envelope and in the interests of obtaining smoothly changing aerodynamic derivatives, it became usual to fix the separation line at a sharp leading edge; there was an

additional advantage in that the vortex provided an 'extra', non-linear component of lift. It was found by Weber^^ and later by Davies^^ that such a sharp-edged wing could produce acceptably low drag up to moderate supersonic Mach numbers if suitably cambered and twisted. The design method used was derived from potential flow theory (and was therefore an 'attached' flow solution) but the drag advantages over a flat wing persisted even when there was a fully developed vortex flow.

Interest in round edged wings lapsed for some time until the start of the 'reusuable' space-shuttle programme. Because of the extremely high speeds (M " 30) at which the orbiter will re-enter the earth's atmosphere, the problems of heat transfer to shaz*p leading edges become insuperable and very rounded leading edges have to be used. The early problems, which led designers

(25)

15

to abandon this type of wing, are suppressed by means of artificial stability augmentation.

Returning to the early work by Stanbrook and Squire^2, we see that for rounded leading edges there is only a very diffuse

boundary between regimes of attached and separated leeside flows, and that this is less well defined than that for thin, sharp-edged wings (figure 21). This is to be expected as the leading-edge radius is an additional unrepresented parameter. It was noted that 'mixed' (i.e. partially separated) flows are common on wings with blunt leading edges; this will also have the effect of increasing the width of the boundary.

The shock detachment boundary is of course given by an = 0, i.e. the shock is detached for all incidences and Mach numbers.

(In practice it is impossible to produce an 'absolutely' sharp leading edge and the shock will be locally detached on any wing. The distinction between sharp and blunt-edged wings becomes

arbitrary and the 'independence principle' (of upper and lower surface flows) not rigorously justifiable. However, it has been found (reference 17) that for moderate amounts of leading-edge blunting the effect on the flow field is felt only over a distance comparable to th6 leading edge radius. A wing will therefore be considered to be 'sharp-edged' if the leading edge radius << £, where £ is"* a characteristic dimension of the wing.)

d) Effect of Reynolds Number

Little systematic work has been conducted to investigate the effect of Reynolds Number on delta wing flow fields, most of the experiments treating only the 'transition free' and 'transition fixed' cases (e.g. references 18 and 19).

The state of the boxindéucy layer on the compression surface has only a minimal effect on the compression surface pressure

distributions because of the favovircüDle pressure gradients existing there. In cases where the compression-svirface shock is detached though, fluid will be spiilled around the leading edges and therefore a turbulent compression-surface boundary layer may cause transition

(26)

Our lack of knowledge concerning the effect of Reynolds Number is reflected for instance in ém inability to predict when shock induced separation will occur in otherwise attached leeside flows. When a vortex flow occurs on the leeside, the theoretical work of Cooke2° leads us to expect a 'saw-tooth' type of transition fz^7nt, different mechanisms causing transition in the attached central region and in the region of strong cross-flow. The main effect of increasing Reynolds Number in this model will be an outboard shift of the suction peaks as the boundary layer becomes turbulent and secondary separation is delayed.

Vortex bursting is also known to be a function of Reynolds Number though there is no theory to predict the conditions under which it will occur. Data for vortex breakdown at low speeds have been obtained by Hummel and Srinivasan^^ who show that the effect of the burst reaching the trailing edge of the wing is to cause a small but sudden change in lift and pitching moment.

Suggestions by Szodruch and Squire^ that vortex burst at supersonic speeds can actually result in positive leeside pressure coefficients were investigated in Part 2.

3.2.1 Thin shock-layer theoiy

In 1963 Messiter^^ published a paper entitled 'Lift of Slender Delta Wings According to Newtonian Theory'. In it he describes a prediction method, later to be referred to as 'thin shock-layer theory', for the forces on thin delta wings at hypersonic speeds. Since then, thin shock-layer theory has been developed by Squire, Hida, Roe, Woods and others into a most versatile and powerful technique. In particular, the range of validity has been extended to include moderate supersonic Mach numbers and low incidences, and the restriction to thin wings (and indeed to conical wings) has

been lifted. Pressure distributions and shock shapes êire accurately predicted and recently Squire has suggested that changes in the

compression surface streamline patterns, calciilated from thin shock layer theory, may provide theoretical justification for the Stanbrook-Squire boundary.

(27)

17

In this chapter we will describe, in chronological order, development of Messiter's theory.

The semi-empiriced Newtonian force law, which predicts a

surface pressure coefficient of the very simple form, Cp = 2 Sin^ o, where o is the local body slope, has been found to give surprisingly good agreement with experiment at hypersonic speeds, though the

physical model used originally to derive the expression seems to be unrealistic. The incident fluid is considered to consist of massy, non-interacting, inelastic particles which, on collision with the body, are assumed to have their normal component of momentum destroyed.

We therefore picture a flow in which the fluid strikes the body and turns abruptly to flow around the body in an infinitesimally thin layer. Although the concept would not have occurred to Newton, we now see that this is tantamount to saying that the shock and body surface are coincident.

The same result may be more credibly derived from the oblique shock relations. In the dual limit M. -»' <», y -»' 1, the shock wave angle becomes equal to the flow deflection angle and the pressure

coefficient behind the shock equal to the Newtonian value.

In 1963 Messiter^^ published a paper in which he set out to iinprove the accuracy of Newtonian theory by assuming that the shock layer was not of infinitesimal thickness but merely 'thin' (i.e. the shock lay close to the body surface). Figure 2 shows the Cartesian coordinate system used, the wing, which is flat, lies in the plane defined by y = 0. For the steady isentropic flow of an ideal gas, we require to satisfy the following equations of motion:

continuity, div p £ = 0

momentum, S*^a •*" C"'-/p)^P = 0 (3.3) entropy, a.V(P/p'') = 0

(28)

together with the shock relationships:

(p(a.n)) = 0

(P + p(£.n)2) = 0 (3.4)

( a X n) = 0

where the larger brackets denote the change in the enclosed quantity across the shock, and n is the unit normal to the shock wave.

Since the shock layer is assumed to be thin, the shock inclination must be approximately equal to the wing incidence, a, and the inverse density ratio across it given by:

It is next required to scale the natural coordinate system so that it will be described by quantities of order unity. For a chord of unit length, the shock stand off distance at the trailing edge ys * etana. The choice of scaling for 3 is less obvious but it is assumed that in the region of interest the aspect ratio of the wing will be of the same order as the Mach angle in the shock layer, which can be shown to be O(e^). The following coordinate 'stretching' is therefore adopted:

X* = X

y* = y/etana z* = z/e*tana

(3.6)

where tana in the third, equation has been introduced to simplify the later algeba.

To estimate the order of magnitude of the independent variables the related solutions for the flow over a swept wing with an attached shock wave is considered and the flow quantities are assumed to have the same order of magnitude everywhere as they have there. The

following equations are derived from the shock relations (equation 3.4) with the assumption that yg— , yg— and P,yp << 1.

(29)

19

U/u v/U

w/u

P-P.

POP P

= Cos a + c(Sin2a/Cos o)u'* + . = e Sin o V* + ....

= e* Sin a w* + ....

= Sin^a + e Sin^a p*

e - - ï - ^ e(2u* + v*^) - e2(l + p*) + .

(3.7)

where u*, v*, w* and p* are the corrections to the Newtonian values.

Substituting equations(3.7) back into the exact eq\;ations of motion and shock relations and retaining only the lowest order terms in e, the following very simple relations are obtained:

Continuity |Xj + ||J = 0

x-mom.

y-mom.

z-mom,

3u* ^ 4 3u* ^ 4 au* _

-i ; 3 r + v * ^ + w * ^ = 0

av* ^ A 3v* ^ * av*

a;jy>> ^* a 7 ^ * *** aiT

i4

aw* ^ o, aw* ^ 4 aw* . _

(3.8) Shock relations u* s V* 8' W* 8

- ay*/ax*

(ay*/ax*) - (ay*/az*)2 - i

S S

- ay*/az*

p* = 2(ay*/ax*) - (ay*/az*)2 - i

(3.9)

As a final simplification, a restriction is made to conical flows, writing,

y s y*/x z = x*/x

(30)

Equations (3.8) and (3.9) then become: Vy + wj = 0 (v - y) ity + (w - z) ug^ = 0 (v - y) >^ + (w - z) v j = - p^ (v - y) ws- • (w - z) w j = 0 (3.11) V - _ au

where urr = — etc.

and Ug = - Yg + zdyg/dz

Vg « ys - zdyg/dz - {dy^/dz)^ - 1

Pg = 2^8 - 2ld78/dï - (d78/dl)2 - 1

(3.12)

The equations (3.11) have two sets of real characteristics given by: z = const. where ( satisfies

(7

-

v^ ii

C s const.

y)

^ +

(w

-

z) ii

=

ay

az

The first of these chcu?acteristics is an unrealistic feature of the equations, implying as it does that disturbances may be transmitted instantaneously across the shock layer. (Hillier euid Woods^*^ point out that in thin shock-layer theory the compressible aspect of the flow is modelled only at the shock. Within the shock-layer the

equations describe a Newton-Busemann type of flow in which interaction between particles is felt only in centrifugal effects normal to the body, i.e. in the z = 0 direction.) The second characteristic path is the streamline and thez^efore has a counterpart in the full equations.

Fixing the constant on the ( characteristicsby setting (YS* Z ) - Z » the solution for equations (3.8) and (3.9) can be

shown to depend only on the cross-flow velocity w(c). For a detached shock wave, w(c) s z^ and the relationship between body slope and shock shape is given by the following integral equation:

(31)

21

dz

>b

= - w(zi5) " zrz •••

body

"^^^^

• 'b {

^^^^ '

^)

ds (3.13)

This is the basic equation for the solution of w((), once it is known the pressure distribution can be determined from equation (3.14).

The boundary conditions of equation (3.13) are

w(o) = 0 w(ft) s 1 + ft

where ft = b/i e^ tan a and is the 'non-dimensional span' (or aspect z^tio)in thin shock-layer coordinates.

The first of the boundary conditions expresses the fact that there is to be zero cross-flow in the plane of symmetry, the second that the singularity in the shock curvature should occur on the normal characteristic through the leading edge (z - Q).

Because ft and z are the only parameters appearing in the expression for the cross-flow velocity (and hence t^ie pressure

distribution and shock shape), ft has the r61e of. similarity parameter in thin shock-layer theory and will be referred to frequently in later pages.

It is not intended to pursue further the solution of equation (3.13). It will be sufficient to note that Messiter computed

solutions for several thin wings and used his results to predict normal force coefficients (with the approximation of zez^ leeside pressure).

The next modification of the theory was due to Hida^^ who included the effect of thickness (diamond and circular arc cross-sections) in the theory. Unfortimately because of the small radius of convergence of the power series chosen, his analytical solutions for pressure distributions gave poor agreement with experiment. In 1967 Squire^** used a more precise nvimerical method to improve the accuracy of the solution.

(32)

For the diamond section wing, the body slope is given in physical coordinates by

® = -

"•az^ 'b

"/.

or in the 'stretched' coordinate system:

Ms-J

'be* s -

b e ' = - C (3.14)

body

This result, with the sign of C changed, also applies to caret wings. The wing thickness appears explicitly in the expression for the pressure at the body surface:

_ _ zb_

P(zb) = 1 - 'ïï^(zb) + 2(Ao - ƒ w(8)d8) + 2zi, w(zb) o

. f- 1 . _ J _ - l i i < ^ b )

T R S ) ^ ^^ ^

2Cft (3.15)

< (ïï(zb)-zb)^^az J ^ia)- sy

which can be evaluated after w(s) has been found from equation

(3.13), subject to the same boundaury conditions as before. Squire's results for pressure distributions show good agreement with e;q>eriment over a wide range of the parameters ft and C and Mach numbers as low as M = 4. The predicted and measured pressures diverge close to the leading edge though, the theory tending to predict too high a pressure in this region.

The most obvious unrealistic feature of thin shock-layer theory is the assumption that the inverse density ratio e, across a basic shock in the leading edge plane, is representative of that across the real shock. The credibility of this assumption diminishes with

increasing wing thickness and decreasing Mach number, i.e. as the real shock moves away from the wing.

In 1974 Squire^^ proposed a method for improving the accuracy of the theory by moving the basic shock away from the plane of the leading edges.

(33)

23

The expressions for pressure and shock shape may be written in functional form:

P(7 , z) - 2Cft + P*(z/ft , ft , C) (3.16)

ys(z) = yt^^/" . n . c ) (3.17)

If the expansion procedxire is carried out about any other basic shock position, the approximate equations are unchanged but new values of ft and C must be used because of the change in e and hence in the coordinate stretching. In addition, because equation (3.15) contains a term in which the shock shape ys is defined relative to the basic shock, the equation must be amended to read:

P(7 .

z)

= 2Cft - ,

^l^^^f^ I

^) + P*(Vft . ft . C) (3.18)

where the basic shock makes angle ^ with the leading-edge plane. If the basic shock is moved through a small angle ^, then the expression for pressure coefficient in the unmodified theory,

Cp = 2 Sin^ a (l + c P*(z/ft , ft)) becomes

Cp = 2 Sin^ a + 2 Sin2(a + (|»)e' P*(z/n«, ft') where the amehded parameters are shown primed.

Only the correction term has been changed, but because e' < e, the range of validity of the theory appears to have been increased.

By careful comparison with experiment. Squire attempted to find the best position for the 'basic shock'. Considering only the detached shock case he investigated three alternatives:

1. Basic shock in leading-edge plane.

2. Basic shock midway between the leading-edge plane and the plane through the calculated shock on the centreline. 3. Basic shock in the plane of the calculated shock on the

(34)

These are respectively termWd the 'unmodified','half-modified' and modified' theories. In the 'half-modified' and

'full-modified' theories, the basic shock position is calculated by an iterative procedure; this converges rapidly as the shock shape is only weakly sensitive to basic shock position, especially if the value of e is small in the first approximation.

The comparison in reference 25 covered the Mach number range 2 < M < 10 and the results can be readily summarised

1. The accuracy of t.s.i.t, can be improved by moving the basic shock.

2. Where the shock is detached use the 'half-modified' theory. 3. Where the shock is attached use the fully modified theory. 4. Close to shock detachment (ft > 2) use the 'full-modified'

theory - but exercise discretion.

From now on the 'prime' will be omitted from the representation for ft and it will be assumed that:

ft = t)/£ e* tan o shock

where a shock is the inclination of the basic shock. The previous definition in which a = a shock is to be seen as a pcirticular case

('unmodified' theory).

In his concluding remarks to reference 25 Squire attempts to explain the sometimes unexpected accuracy of thin shock-layer theory:-"It is interesting to find that for flat wings the initial movement of the shock below the wing produces only small changes in the surface pressure distribution but, particularly at low incidence, even a small movement of the basic shock can produce a marked reduction in the value of the inverse density ratio e. This suggests that the real governing parcuneter is the inverse density-ratio across the calculated shock wave rather than across the

arbitrarily chosen basic shock position". For iterative procedures, in which the position of the calculated shock wave influences that of the basic shock, this point may well be valid but for the

(35)

25

beginning of the paper, this explanation does not apply and an

expansion in a power series in e appears hard to justify, though an investigation of the higher order terms may reveal that their

coefficients are always small.

For the delta wing tested at Cranfield, the shock is detached for all incidences and therefore the 'half-modified' theory would be expected to give the best results. Figure 22 shows the variation of c with o for this wing in the three versions of the theory.

3.2.2 Extension to the prediction of leeside flows

Recently, Shanbhag^^ has written a program for the solution of equation (3.13) for the cross-flow velocity w(ft , C ) . The numerical method used to solve the equation is different from that used previously by Squire and very much quicker. Considering first, the case C s 0

(thin wings), the solutions for w are shown in figure 23, together with the coz^esponding streamline patterns and pressure distributions. The striking feature is that at ft = ft = 0*5 the streamline pattern appears to change abruptly from a situation in which there is only a single attachment line at the centreline (figure 23a) to one in which there are three attachment lines (figure 23b). In 23a, the conical streamlines are all divergent; in 23b, the conical streamlines are divergent outboard of the outer attachment lines and convergent inboazHl of them. An inspection of figvire 23c suggests that the change in streamline pattern may occur as a result of a singularity when part of the solution becomes tangential to w(s) = s. The mathematical reason for the jump is not really cleêur however and in the real flow the movement of the attachment lines would be expected to occur over a finite incidence range.

In the case of thick wings. Squire^ was only able to compute solutions for w coz^esponding to the single attachment-line type of flow. The solutions lying êüaove the line w(s) = s were well

behaved but no method could be discovered to produce solutions lying below the line. Hence it was inferred that, by analogy with the thin wing behaviour, the flow pattern changed to the three attachment-line type when the solutions crossed this line. The maximum value of ft

(36)

for which the solutions always lie above w(s) = s has been found to be closely given by:

ft = 0*5 + 1*2 C

The curve ft = ft is plotted in figure 20 and shows it heading encouragingly for the Stanbrook-Squire boundary. The two curves do not intersect because of the requirement e < 1, which limits us to high values of incidence when ft is small. We are apparently presented with an attractive model in which leading-edge separation occurs because of a sudden increase in the outflow around the leading edges. This outflow is the result of a change in flow pattern on the compression surface which may be predicted by thin shock-layer theory. The available data for thin wings (e.g. references 27 and 28) tends to support the theory that leading-edge separation occurs for a particular value of ft, though not precisely ft = 0*5. The outboard attachment lines have been observed (again on thin wings) by Bashkin^^ although unfortunately he does not relate their movement to the start of leading-edge separation.

The data for thick wings are extremely sparse and the present investigation cu?ose from one of the suggestions for 'further work' in reference 3. Figure 24 shows the locus of ft = ft for the C.I.T. model together with the exact shock detachment boundary and the Stanbz^ook-Squire boundary. At M = 2*5, ft = ft for a = 10°.

3.3 Tunnel Investigation of a Particular Delta Wing 3.3.1 Aims of the Investigation

The objectives of the experimental investigation were twofold:-i) To produce a delta wing model which could be tested over

a wide range of incidences (0 < a < 50°) without serious support interference. The model also had to be capable of being extensively and accurately pressure plotted. ii) To investigate the boundary between regimes of attached

and separated leeside flows and in particular to relate the results to the movement of the attachment lines on the compression surface.

(37)

27

3.3.2 The Models

Due to the difficulty in accommodating the pressure tubes, two models were constructed; Model 1 with pressure tappings in the flat expansion surface and Model 2 with tappings in the 'V' compression surface. The models were otherwise identical (see figure 25). The pressure tubes were led away through a small diameter hollow sting. This method of construction was adopted because earlier work by the Author^° had shown that a combination of large sting and multiple

pressure tubes in the base region could profoundly affect the leeside flow field. A photograph of the models is shown in figure 26. The surface pressures were measured by means of 2mm diameter hypodermic tube recessed into the model surface and rendered flush using Araldite. Two stings were constructed, one straight and the other bent through 25 degrees. In conjunction with the tunnel incidence adjustment of 30 degrees, this covered the required range with a useful 5 degree overlap to check for possible sting interference. To ensure alignment, the stings were fixed permanently into the model bases. The models themselves were in two parts, split just forward of the trailing edge to allow the sting assemblies to be interchanged; the two halves were secured by dowels and screws. This arrangement had the additional merit of providing a very 'clean' base area.

After machining, the models were set up on a surface tcdile and carefully.,measured. Slight 'bowing' was found to have occurred as a result of the installation of the pressure tubes. The table below shows the deviation from design of the two models at the centreline of the flat face. The datum is a line through the tip and parallel with the sting. In both cases the flat surfaces were slightly concave. *fl Model 1 Model 2 A y mm 0 0 0 0*2 0*40 0 . 3 1 0 . 4 0*60 0*47 0*6 0*70 0*57 0 . 8 0-75 0 . 6 0

1.0

1

0-75 0 . 6 0

The bend is greatest near the tip, but a line drawn from tip to

(38)

No attempt to allow for this is made in the reduction of the data; all quoted incidences refer to the sting inclination.

3.3.3 Test Prodecure a) Pressure Plotting

The sxirface pressure tubes were initially drilled 2mm from the leading edges to produce static pressure tappings. This represented the closest approach to the leading edges with this tube layout and was equivalent to about 95% of the semispan at 90% chord. The hole diameter was 0.25mm. The eight or nine ports per model were scanned sequentially by a Statham' 0-10 psia transducer using an NPL pressure scanning switch mounted external to the tunnel. Free stream total and static pressures were also scanned and the output fed via a Dynamco data logger to a teletypewriter. The results were reduced to the nondimensional ratio P/pso.

Figvire 27 shows a general view of the tunnel working section and pressure measuring system.

On subsequent runs the pressure tubes were drilled successively nearer the model centreline and the outboéurd holes were plugged with beeswax.

b) Schlieren Photographs

A conveiTtional single-pass schlieren system was used. The knife-edge was aligned with the free-stream for all the photographs shown here. Other orientations were investigated but provided little extra information. Poloucoid film was used initially but was not found to give consistent picture quality. Ilford blue-sensitive plates (type LN) were used subsequently with good results. A

continuous mercury vapour source was used and exposure times of •v» 1/200 S were provided by the camera shutter.

c) Vapour Screen Technique

The static temperature in the working section of a supersonic tunnel is, as a rule, very much lower than ambient temperature. It is therefore usual to dry the air in the tunnel to prevent a mist of condensation forming in the working section. If illuminated by a

Cytaty

Powiązane dokumenty

The research and design work carried out by nu­ merous institutes as well as by engineering and technological offices was accompanied by a campaign led by

Niewiele reliktów miało pozostać ze starej świątyni w Korytnicy Łaskarzewskiej 67. znajdujemy też wzmiankę, „że projekt powiększenia kościoła był - lecz

W rozważanej przez nas sytuacji w praktyce wynika stanowczy zakaz podawania takich przyczyn odmowy przyjęcia obrony, które stanowiłyby swoisty „przedwyrok” w

Aktywność fizyczna, sposób odżywiania się i zawartość tkanki tłusz- czowej u chłopców trenujących i nietrenujących w wieku 14 – 15 lat.. Physical activity, dietary habits

Obtained SchRs 4000/50 excavator operation technology using “the system” as well as range and effect of practical usage of designing and immediate planning of bucket ladder

Dane dotyczące fal upałów obejmują daty wystąpienia, czas trwania oraz charakterystykę meteorologiczną wszystkich fal upałów, jakie wystąpiły na 16 stacjach w Polsce,

ding is nogal groot. Dit is vooral te zien in de figuur voor de treksterkte waar twee proefplaten per hoek beschikbaar waren. Voor de E bepaling werd siecht 4n serie

For the sequence learning, the following features were used: The end-effector position in the cylindrical frame and the Cartesian frame as well as the data from the force sensor at