Top Banner
National Aeronautics and Space Administration Langley Research Center Hampton, Virginia 23681-0001 NASA Technical Paper 3559 Aerodynamic Characteristics of Two Waverider-Derived Hypersonic Cruise Configurations Charles E. Cockrell, Jr. and Lawrence D. Huebner Langley Research Center • Hampton, Virginia Dennis B. Finley Lockheed-Fort Worth Company • Fort Worth, Texas July 1996
76

Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

Jun 25, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

National Aeronautics and Space AdministrationLangley Research Center • Hampton, Virginia 23681-0001

NASA Technical Paper 3559

Aerodynamic Characteristics of TwoWaverider-Derived Hypersonic CruiseConfigurationsCharles E. Cockrell, Jr. and Lawrence D. HuebnerLangley Research Center • Hampton, Virginia

Dennis B. FinleyLockheed-Fort Worth Company • Fort Worth, Texas

July 1996

Page 2: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

Printed copies available from the following:

NASA Center for AeroSpace Information National Technical Information Service (NTIS)800 Elkridge Landing Road 5285 Port Royal RoadLinthicum Heights, MD 21090-2934 Springfield, VA 22161-2171(301) 621-0390 (703) 487-4650

Available electronically at the following URL address: http://techreports.larc.nasa.gov/ltrs/ltrs.html

Page 3: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

1. Introduction

A waverider is any shape designed such that the bowshock generated by the shape is perfectly attached alongthe outer leading edge at the design flight condition. Thewaverider design method leads to several potentialadvantages over conventional non-waverider hypersonicconcepts. The attached leading-edge shock wave con-fines the high-pressure region to the lower surface andresults in high lift-drag ratios. Several design predictionssuggest that waveriders may offer an aerodynamic per-formance advantage in terms of higher lift-drag ratiosover non-waverider hypersonic concepts (refs. 1 and 2).In addition, the flow field below the waverider bottomsurface is uniform and, in the case of waveriders derivedfrom axisymmetric flow fields, there is little or no cross-flow in this region, making these shapes attractive candi-dates for engine integration. These advantages have ledto interest in using waverider shapes for the forebodygeometries of hypersonic airbreathing engine-integratedairframes. Waveriders have been considered for varioustypes of missions including hypersonic cruise vehicles,single-stage-to-orbit vehicles, airbreathing hypersonicmissiles, and various space-based applications (ref. 3).

The purpose of the current study is to examine theaerodynamic characteristics of two waverider-derivedhypersonic cruise vehicles. No experimental data cur-rently exist that address the integration of realistic vehi-cle components with waverider shapes. Therefore, theobjectives of this study were threefold. The first was tocreate an experimental and computational database forwaverider-derived configurations. The second was toexamine the effects of individual vehicle components onpure waverider performance and to determine the differ-ences in aerodynamic characteristics that result from

integrating all vehicle components. The final objectivewas to evaluate the controllability of each of the fullyintegrated vehicles and the effectiveness of the control-surface design. These objectives were accomplishedusing results from wind-tunnel testing and a limited num-ber of computational fluid dynamics (CFD) solutions.The CFD predictions were obtained for the pure wave-rider shapes only and provide comparisons with experi-mental data and design-code predictions. Two wind-tunnel models were designed that integrate canopies,engine packages, and control surfaces with two Mach 4.0pure waverider shapes. The models were tested in theLangley Unitary Plan Wind Tunnel (UPWT) at NASALangley Research Center.

This report describes the waverider aerodynamicdesign code used and discusses the method used in thedevelopment of the wind-tunnel models. The details ofthe experimental study are then presented as well as thecomputational method used to obtain the CFD predic-tions. The results are analyzed in three sections. First, theresults of the pure waverider shapes without integratedvehicle components are presented. These results includeflow-field characteristics from CFD solutions and experi-mental flow-visualization data as well as aerodynamiccharacteristics from the experiment and CFD predictions.Second, the experimental results of adding aircraft com-ponents to the pure waverider shapes are presented. Theeffects of the canopy, engine components, and controlsurface additions on aerodynamic performance andstability are examined. Finally, the aerodynamic charac-teristics of the fully integrated waverider-derived config-urations are examined and compared with those of thepure waverider shapes. Control-surface effectiveness isalso addressed in this section.

Abstract

An evaluation was made of the effects of integrating the required aircraft compo-nents with hypersonic high-lift configurations known as waveriders to create hyper-sonic cruise vehicles. Previous studies suggest that waveriders offer advantages inaerodynamic performance and propulsion/airframe integration (PAI) characteristicsover conventional non-waverider hypersonic shapes. A wind-tunnel model was devel-oped that integrates vehicle components, including canopies, engine components, andcontrol surfaces, with two pure waverider shapes, both conical-flow-derived wave-riders for a design Mach number of 4.0. Experimental data and limited computationalfluid dynamics (CFD) solutions were obtained over a Mach number range of 1.6to 4.63. The experimental data show the component build-up effects and the aero-dynamic characteristics of the fully integrated configurations, including control sur-face effectiveness. The aerodynamic performance of the fully integrated configura-tions is not comparable to that of the pure waverider shapes, but is comparable topreviously tested hypersonic vehicle models. Both configurations exhibit good lateral-directional stability characteristics.

Page 4: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

2

2. Symbols

B.L. buttline of model (distance from centerline inspanwise direction), in.

CD drag coefficient

Cl rolling-moment coefficient

Clβ rolling-moment derivative,

CL lift coefficient

CM pitching-moment coefficient

Cn yawing-moment coefficient

Cnβ yawing-moment derivative,

moment reference length, in.

L/D lift-drag ratio,

M Mach number

M.S. model station (distance from nose in stream-wise direction), in.

P pressure, lbf/ft2

Pss roll rate, deg/sec

Re Reynolds number

Sref planform area, ft2

u velocity component, ft/sec

V total volume, ft3; velocity, ft/sec

Veff volumetric efficiency,

W.L. waterline of model (distance from zero refer-ence in vertical direction), in.

Xc.g. moment reference center location

X, Y, Z Cartesian coordinates, in.

y+ inner law variable

α angle of attack, deg

β sideslip angle, deg

δA angle of aileron deflection (trailing edge downpositive), deg

δE angle of elevon deflection (trailing edge downpositive), deg

∆ζ distance from solid boundary to first cellcenter, in.

µ viscosity coefficient, lbf-sec/ft2

ξ, η, ζ computational coordinates

ρ density, lbm/ft3

Subscripts:

c conditions at first cell center next to solidboundaries

c.g. center-of-gravity location

∞ free-stream conditions

3. Configuration Design and ModelDevelopment

3.1. Waverider Design Method

A specific waverider shape is uniquely defined byfree-stream conditions, the type of generating flow-fieldbody, and a leading-edge definition (ref. 1). The shapesof the upper and lower surfaces of the configuration fol-low from these parameters. The free-stream conditions,including Mach number and Reynolds number or alti-tude, are selected based on mission criteria. The designmethod used in this study involves a specific designpoint. The generating flow-field body is used to definethe shock shape upon which the leading edge of thewaverider is constructed. Although any arbitrary body insupersonic or hypersonic flow can be used as a generat-ing flow-field body, this study focuses specifically on theclass of conical-flow-derived waveriders, in which thegenerating flow-field body is a right circular cone insupersonic or hypersonic flow. At the outset of thisresearch effort, this option was the best available for theapplication of interest. Other possible generating flowfields include osculating cone flow fields (ref.4), hybridcone-wedge generated flow fields (ref.5), and inclinedcircular and elliptic conical flow fields (ref.6). Thelength of the generating cone, length of the waverider,and semiapex angle of the cone are specified by thedesigner. The selection of these parameters can signifi-cantly affect the shape of the waverider generated as wellas the aerodynamic performance of the configuration.Figure 1 illustrates the design of a conical-flow-derivedwaverider. The planform shape, or leading edge, isdefined on the shock wave produced by the cone. Thelower surface of the configuration is defined by tracingstreamlines from the leading edge to the base of the cone.The result is that the lower compression surface is astream surface behind the conical shock wave. The con-figurations studied here have an upper surface that isdesigned as a constant free-stream pressure surface.However, other techniques may be used, such as shapingthe upper surface as an expansion or compressionsurface. The conical flow field, defined behind theshock wave, exists only below the lower surface ofthe waverider.

∂Cl

∂β---------

∂Cn

∂β----------

c

CL

CD-------

V2 3⁄

Sref------------

Page 5: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

3

The resulting configuration offers two possibleadvantages over non-waverider hypersonic configura-tions. The first is a potential aerodynamic performanceadvantage (refs. 1, 2, and 7). Theoretically, the shockwave is perfectly attached along the outer leading edge atthe design Mach number. The result is that the high-pressure region behind the shock wave is confined to thelower surface, and no flow spillage from the lower sur-face to the upper surface occurs. The maximum lift-dragratios this method produces promise to exceed those ofexisting hypersonic configurations. Figure 2, taken fromreference 2, shows the traditional “L/D barrier” in thesupersonic/hypersonic regime for conventional vehicles.This correlation is empirical, based on actual flight vehi-cle experience at subsonic and low supersonic speeds andextrapolated to hypersonic Mach numbers (ref. 7). Thesymbols in figure 2 represent predictions for a variety ofconical-flow-derived waverider shapes generated usingthe current method, which is described in detail in refer-ence 2. The waverider shapes represented here are onlythe forward portions of possible hypersonic configura-tions and therefore are not realistic vehicles. The predic-tions shown assume that the configuration has zero basedrag in order to remove the effect of the blunt base,which will be eliminated in a fully integrated vehicle,and show only the performance of the forward portion ofsuch a vehicle. In other words, the predictions assumethat free-stream static pressure acts at the base, making adirect comparison of the lift-drag ratios for waveridersand those of existing supersonic/hypersonic configura-tions difficult. Furthermore, the waveriders representedhere do not have levels of volumetric efficiency compa-rable to those of the vehicles used in theL/D barrier cal-culation and may not have been obtained at similarflight-scaled Reynolds numbers. Although the lift-dragratios of a fully integrated waverider configuration withthe blunt base closed would likely be lower than thosefor the pure waverider shape, these predictions suggestthat waveriders may offer an aerodynamic performanceadvantage over non-waverider vehicle concepts. Anotheradvantage of axisymmetric waverider flow fields is thatthe lower surface flow field is uniform, and there is pureconical flow in this region for a perfectly attached shockwave. Therefore, a known uniform flow field can bedelivered to scramjet engine modules on the lower sur-face, providing a benefit in propulsion/airframe integra-tion (PAI) (ref. 8). The osculating cone and cone-wedgeconcepts mentioned previously may provide an evengreater benefit over conical-flow-derived waveriders(refs. 4 and 5). The aerodynamic performance and PAIbenefits suggested in previous research efforts have gen-erated interest in using waveriders for various hypersonicvehicle designs.

The design code used in this study is the (Universityof) Maryland Axisymmetric Waverider Program(MAXWARP) (refs. 1, 2, and 9). The MAXWARP codeis an inviscid design method that includes an estimate forskin friction in the design process. Various volumetricconstraints may also be imposed by the user in order toproduce waveriders with desirable structural characteris-tics and component packaging. These constraints includeaspect ratio, slenderness ratio, and total volume. Forthe case of conical-flow-derived waveriders, the Taylor-Maccoll equation, which describes the flow field behinda conical shock wave (ref. 10), is integrated using afourth-order Runge-Kutta method to compute the invis-cid conical flow field behind the shock wave. The conesemiapex angle and length of the flow-field generatingbody are specified by the user along with free-streamconditions. The code starts with an initial leading-edgedefinition on the conical shock wave and creates awaverider shape from this initial leading edge. The pres-sure distributions on the surface of the configuration areintegrated to calculate lift and drag coefficients. An esti-mate for skin friction is also included so that force coeffi-cient predictions include both inviscid and viscouseffects. This estimate is based on the reference tempera-ture method, which is described in reference 11. Theeffect is to generate shapes for which wetted surface areais minimized to reduce skin friction drag. The code usesa simplex optimization routine (ref.1) to optimizewaveriders for a given figure of merit: maximum lift-drag ratio or minimum drag. More recent versions of thecode allow the user to construct various other objectivefunctions. At each iteration in the optimization process,an updated leading-edge definition is used to generate anew waverider shape that progresses toward the desiredfigure of merit. This process continues over a number ofiterations until the optimum shape is found without viola-tion of any of the user-specified volumetric constraints.

3.2. Waverider Shape Description

The pure waverider shapes used in this study, whichdefine the forward portions of the waverider-derivedvehicles, were designed using the MAXWARP designcode. Free-stream conditions and optimization parame-ters were chosen based on the applicability of this studyto a hypersonic cruise vehicle, with available ground-based test facility limitations taken into account. Thedesign free-stream Mach number was 4.0 and the designReynolds number was 2.0× 106 per foot. Although thespecific cruise Mach number for this type of vehiclewould be higher, Mach 4.0 was selected as the designpoint based on the limitations of the UPWT and therange of data desired. The Mach number range of this

Page 6: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

4

facility is 1.47 to 4.63. A design point of Mach 4.0 wouldpermit the validation of the waverider concept at thedesign Mach number and also allow for the determina-tion of aerodynamic characteristics at off-design Machnumbers. The use of endothermic fuels on this vehicleclass is expected to drive the selection of cruise Machnumber to approximately 5.0 to 5.5. No significant dif-ferences in the flow physics are expected between theultimate design Mach number and the Mach numberrange investigated in this study. The Reynolds numberchosen is based on nominal facility operating conditionsin the UPWT and is not representative of a flight cruisealtitude. The configuration was optimized for maximumlift-drag ratio at the design point because this quantity ismore appropriate than minimum drag as a hypersoniccruise performance parameter.

A fully turbulent boundary layer and a wall tempera-ture of 585°R were specified in the design. This walltemperature was selected based on previous experimentaldata from models tested in the UPWT. It is not likely thatfully laminar conditions could be maintained in experi-mental testing at the conditions of interest, and transitionis difficult to predict. Fully turbulent conditions can beachieved and maintained by the application of boundary-layer transition grit to the model surface.

Two different pure waverider shapes were developedfor this study. The first is referred to as the “straight-wing” shape and was designed using the MAXWARPoptimization routine. The second, referred to as the“cranked-wing” shape, was created by adjusting the lead-ing edge of the straight-wing waverider to create acurved wingtip shape that had increased aspect ratio butstill maintained shock attachment along the outer leadingedge at the design free-stream condition. The term“cranked” in this context refers to a wing shape in whichthe sweep angle not only changes but also exhibits alarge outboard dihedral angle in the plane of the base.The cranked-wing shape was designed to provideimprovements in subsonic aerodynamic performance(because of increased aspect ratio) and in lateral-directional stability (because of dihedral effect) whilemaintaining high performance in the supersonic/hypersonic regime.

Three primary design criteria were used to select thebest waverider shape designs for this application. First,the maximum lift-drag ratio was chosen to be as high aspossible while not violating other design guidelines. Thiscriterion drives the selection of the cone semiapex anglefor the generating flow field. A value of 8.1° wasselected for this application. Second, the volumetric effi-ciency (V2/3/Sref) was chosen to be as high as possible.An inverse relationship exists between the volumetricefficiency and the maximum lift-drag ratio for a given set

of free-stream conditions. Therefore, an attempt wasmade to increase the volumetric efficiency as much aspossible while accepting a minimum penalty in maxi-mum lift-drag ratio. Finally, a configuration with a flat orslightly convex bottom surface in the cross section wasdesired for ease in propulsion systems integration. Inaddition to these three primary design guidelines, a con-figuration free of substantial curvature over most of thecross section was also desired to provide for the inclusionof an internal spar in an actual aircraft. Furthermore, thetarget value of span-to-length ratio was 0.8. Informationfrom previous studies shows that larger span (higheraspect ratios) waveriders provide higher lift-drag ratiosbut are more difficult to integrate as a full waverider-based vehicle (ref. 12).

A three-view drawing and an oblique view of thestraight-wing pure theoretical waverider shape generatedby the design code are shown in figure 3. Table 1 sum-marizes the characteristics of this shape. The span-to-length ratio is 0.83. The lower surface of the straight-wing configuration has a slight convex curvature thatfacilitates integration of the propulsion system. Thelength selected for the waverider configuration was24.0in. based on the size of the test section in theUPWT. The length of the generating cone was selectedto fix the location of the waverider leading edge on theconical shock wave to achieve the design criteria notedpreviously (48.0 in. for this application). A selection ofdifferent locations on the conical shock wave wouldresult in waveriders with much different geometric char-acteristics and may result in the generation of unrealisticshapes that could not be integrated into vehicles. Thevolumetric efficiency,Veff, of this configuration is 0.11with a predicted maximum lift-drag ratio of 6.9.

A three-view drawing and an oblique view of thecranked-wing pure theoretical waverider shape generatedby the design code are shown in figure 4. The crankedleading edge still lies on the same conical shock waveproduced by the generating cone used to design thestraight-wing waverider. The characteristics of thecranked-wing waverider shape are summarized intable2. The span-to-length ratio is 0.96, which representsan approximately 16 percent increase in aspect ratio.This increase in aspect ratio should improve the subsonicaerodynamic performance over the straight-wing wave-rider while maintaining the structural characteristicsof the straight-wing waverider near the centerline ofthe configuration. The volumetric efficiency of thisconfiguration is 0.108 with a calculated maximum lift-drag ratio of 6.7. This configuration represents only aslight decrease of both parameters from the straight-wingwaverider. The slight convex curvature of the bottomsurface is maintained toward the centerline of themodel. The dihedral angle of the aft cranked section is

Page 7: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

5

approximately 28° when measured from the centerline ofthis section.

The values for maximum lift-drag ratio given are forthe pure waverider shapes only. The waveriders weresubsequently altered to close the blunt base and add con-trol surfaces. The predictions assume that free-streamstatic pressure are acting at the base of the unaltered purewaverider shape, so that only forebody drag values areincluded in the performance predictions. As will beshown later, the incorporation of aftbody closure is a sig-nificant issue in hypersonic vehicle development.

3.3. Wind-Tunnel Model Designs

Two slight modifications to the design-code shapeswere implemented in the wind-tunnel model design inorder to accommodate model support hardware and addi-tional vehicle components. A smooth ogive-cylindricalfairing was blended on to the upper surface of the purewaverider shapes to accommodate the sting and balancenecessary to measure the aerodynamic loads on themodel during testing. This volume was added to theupper surface rather than the lower surface because pre-vious research indicates that modifications to the lowersurface have an affect on the PAI characteristics of thewaverider (ref. 13).Figures 5 and 6 show tunnel installa-tion photographs of the straight-wing and cranked-wingpure waverider models with the upper surface fairing.The lower surface of the theoretical waverider shape wasmodified slightly by creating an inboard expansion sur-face with an angle of approximately 10°, beginningapproximately 22 in. aft of the nose of the configurationand measuring approximately 3.5 in. in the spanwisedirection. The lower surfaces follow the waverider theo-retical stream surface up to this point. This modificationwas made in order to facilitate the integration of enginecomponents and to reduce the closure angle necessary forcontrol surfaces. Figure 7 shows a photograph of thelower surface of the cranked-wing waverider with theexpansion on the aft end of this surface.

A realistic canopy was designed for the waverider-based configuration. The canopy was provided with fac-eted surfaces to resemble the canopy for a hypersonicvehicle. The aft portion of the canopy was designed toblend with the cylindrical fairing on the upper surfacediscussed previously. Figures 5 and 6 show the purewaverider models with the ogive-cylindrical fairingattached (i.e., canopy-off configuration). Figures 8 and 9show the model with the faceted canopy attached.

The engine package for this configuration included acompression ramp, a non-flow-through engine modulewith side walls, and a nozzle/expansion ramp. Theengine-package-on configuration provided an indicationof the effect of modifying the theoretical waverider lower

surface to integrate some type of engine system and isnot intended to be a realistic propulsion simulation. Theinlet capture area, expansion ramp turning angle, andnozzle exit area were designed for full-scale Mach 4.0conditions. The compression surface shown in figure 8 isrequired for additional precompression of the flow enter-ing the inlet. The non-flow-through configurationattempts to model the external cowl drag present on arealistic flow-through nacelle, but does not have theassociated internal drag. Two different nozzle/expansionramps were designed for the model. The first was usedwith the pure waverider configurations with the nacelleattached and the second was used with configurationsthat had control surfaces attached. These nozzles arereferred to as the “short” and “long” nozzles, respec-tively (figs. 10(a) and 10(b)). Identical nozzles with staticpressure taps were also fabricated in order to obtain sur-face pressure measurements on the nozzle. The non-instrumented ramps were used for force and momentruns.

Control surfaces were provided to examine theireffects on waverider aerodynamic performance as well asthe effectiveness of the control concept. The control sur-faces were sized based on control-volume trends fromsupersonic fighter aircraft to extend and close the bluntbase of the configurations. Elevon deflections of 0°, pos-itive 20° (trailing edge down), and negative 20° (trailingedge up) were incorporated. A set of outboard aileronshaving the same three deflection angles was designed forthe straight wing. Because of the curved surface of thecranked wing and the small thickness of the outboardleading edge, the set of ailerons for the cranked-wingconfiguration consisted of an inboard aileron, whichremained fixed at 0°, and a set of outboard ailerons,which were deflected at 0° and±20°. A vertical tail sur-face was also designed in order to augment directionalstability. Figures 8 and 9 show photographs of the modelcomponents with the various control surfaces. Figure 11shows three-view drawings of the elevons, straight-wingailerons, cranked-wing inboard ailerons, and cranked-wing outboard ailerons. This figure indicates the perti-nent dimensions and shows the hinge-line locations foreach control surface.

The model design allowed for testing of the straight-wing and cranked-wing pure waverider models, whichare defined as configurations with no engine componentsor control surfaces. A configuration build up of thewaverider models with different vehicle componentscould also be tested up to and including the fully inte-grated waverider-derived configurations, which aredefined as configurations with engine components, con-trol surfaces, and the canopy. Table 3 shows the pertinentmodel geometry for each configuration tested. Figure 12

Page 8: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

6

shows a three-view drawing of the fully integratedconfigurations.

4. Experimental Method

The facility used in this study was the UPWT atNASA Langley Research Center. The UPWT is a closed-circuit, continuous-flow pressure tunnel with two 4- by4- by 7-ft test sections, which were both used in thisstudy. The Mach number range of the facility is 1.47to 4.63, with a range in the low Mach number test sectionof 1.47 to 2.86 and a range in the high Mach number testsection from 2.30 to 4.63. Continuous variation of Machnumber is achieved by using asymmetric sliding blocknozzles to vary the nozzle throat-to-test-section arearatio. The Reynolds number range of this facility is0.5× 106 to 8.0× 106 per foot. However, the nominalReynolds number for most tests is 2.0× 106 per foot.A detailed description of the UPWT can be found inreference 14.

The configurations tested ranged from the straightand cranked pure waverider models to the fully inte-grated waverider-derived vehicles. The test configura-tions were chosen to show pure waverider performance;to isolate the effects on waverider aerodynamic perfor-mance of the canopy, engine package, and control sur-faces; and to show the aerodynamic performance andstability characteristics of the fully integrated configura-tions. Only the cranked-wing configurations were testedin the low Mach number test section. The data were cor-rected for flow angularity in the test sections. Calibrationdata for the UPWT shows that the flow in both test sec-tions has an upflow angle generally within 0.5° of thetunnel centerline (ref. 14). In each run, either six-component force and moment data, nozzle pressure data,or vapor-screen photographs were obtained. Schlierenphotographs were taken during the force and momentruns.

The test conditions were chosen to investigate theaerodynamic performance and stability of each configu-ration at both the design Mach number and at off-designMach numbers. Data were obtained at Mach numbers of2.3, 4.0, and 4.63 for all configurations studied and, addi-tionally, at Mach numbers of 3.5 and 4.2 for some con-figurations. Data for the cranked-wing configurationswere also obtained at Mach numbers of 1.6, 1.8, and 2.0.The free-stream Reynolds number for most runs was2.0× 106 per foot. Some runs were made at Reynoldsnumbers of 1.5× 106 per foot and 3.0× 106 per foot inorder to investigate the effects at off-design Reynoldsnumbers. The angle-of-attack range studied was−6° to10o at fixed sideslip angles of 0° and 3°. Data wereobtained over a sideslip angle range of−5° to 5° for thefirst configuration run in each test section in order to ver-

ify that yawing and rolling moment values are linear overthis range (ref. 15). Based on these results, stabilityderivatives were calculated from data obtained at the twofixed sideslip angles.

The data obtained from the wind-tunnel tests includesix-component force and moment data, static pressurereadings on the blunt base of the model, static pressuredata on the nozzle surfaces, and flow-visualization data.The balance used in this case was the NASA-LaRC-designated UT-50-B balance, which is a six-componentstrain gauge balance. Unless otherwise noted, themoment reference center for all configurations waslocated 16.623 in. aft of the nose. A total of 11 5-psipressure transducers were used to measure the staticpressure along the blunt base of the configurations and inthe cavity surrounding the sting. Integrated areas wereassigned to each tap or averaged group of taps and usedto calculate the base axial force. All of the force data pre-sented is corrected to assume free-stream static pressureacting at the base. This procedure is carried out so thatthe data may be presented showing only the upper andlower surface lift and drag values and eliminating theeffect of the blunt base. This procedure is necessarybecause the base will be eliminated in any realisticwaverider-derived configuration. The method of assum-ing free-stream pressure at the base is consistent with thedesign-code method and with previous studies showingpredictions for waverider aerodynamic performance(refs. 2, 9, and 16). Details on the procedure used areincluded in reference 15. For configurations with bothengines and control surfaces, only two base and twochamber pressures were measured. A 32-port, 5-psiexternal electronically scanned pressure (ESP) modulewas used to measure the static pressure on the nozzlesurface for four runs. Figure 10 shows the locations ofpressure taps on the nozzle surfaces for the short andlong nozzles. Recall that the short nozzle is used withconfigurations having no control surfaces and the longnozzle is used with configurations with control surfaces.A total of 12 pressure taps were located on the short noz-zle and 24 pressure taps were located on the long nozzle.The data are used to correct the nozzle surface pressuresto assume free-stream static pressure acting on these sur-faces for some configurations.

Schlieren and laser-vapor-screen photographs weretaken in order to examine flow-field features includ-ing the shock attachment characteristics for variousconfigurations. For the vapor-screen runs, the laser waspositioned outside of the test section window and thelight sheet was projected across the model surface in thespanwise direction, illuminating one cross section ata time. The camera was mounted inside of the test sec-tion above and behind the model. This setup gives a

Page 9: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

7

cross-sectional view of the waverider flow field in thevapor-screen photographs.

The accuracy of the UT-50-B balance, based on aMay 1993 calibration, is 0.5 percent of full scale for eachcomponent to within 95-percent confidence. The full-scale load limits were 600 lbf normal, 40 lbf axial,1500in-lbf pitching moment, 400 in-lbf rolling moment,800 in-lbf yawing moment, and 300 lbf side force. As anexample, using the method of root-mean-squares sum-mation to combine independent error sources, these lim-its correspond to a range of uncertainty in lift coefficientof 0.0053 atα = 0° to 0.0054 atα = 10° and an uncer-tainty range in drag coefficient of 0.00036 atα = 0°to 0.001 atα = 10° for theM∞ = 4.0 andRe∞ = 2.0× 106

per foot condition. The repeatability of measurementswas observed to be better than these uncertainties. There-fore, differences less than the indicated ranges for com-parisons with data from different configurations in thesame test, could be considered significant. However,comparisons between independent measurements areonly good to within the quoted uncertainty ranges. Tran-sition grit (no. 60 size sand grit in the low Mach numbertests section and no. 30 size grit in the high Machnumbertest section) was applied in a 0.1-in-wide strip to themodel upper and lower surfaces along the outboard lead-ing edge at a location approximately 0.4 in. from theleading edge in the streamwise direction. These proce-dures were established for models tested in the UPWTbased on unpublished transition experiments conductedin the UPWT and the methods of references 17 to 19.

5. Computational Method

Computational grids were developed for each of thepure waverider configurations by first developing anumerical surface description and then creating 3-Dvolume grids. Numerical surface descriptions of thestraight-wing and cranked-wing wind-tunnel modelswere obtained from computer-aided design (CAD)descriptions of the model parts. Three-dimensional vol-ume grids were created for each configuration using theGRIDGEN software package, which uses algebraictransfinite interpolation methods with elliptic interiorpoint refinement (ref. 20). Only the pure waveridershapes with no integrated vehicle components were mod-eled for the CFD analysis.

The computational grids for each of the two purewaverider shapes model only half of the configurationbecause each is symmetric about the centerline. The gridorientation is shown in figure 13. Theξ-computationaldirection runs from the nose of the configuration to thebase in the streamwise direction. Theη-computationaldirection begins at the upper centerline and wraps aroundthe leading edge, ending at the lower centerline. The

ζ-computational direction runs from the surface of theconfiguration to the outer boundary. The grids for eachof the two pure waverider shapes contained 91points inthe ξ direction, 111 points in theη direction, and91points in theζ direction. Blunt leading edges weremodeled for each configuration in order to provide a bet-ter comparison with experimental data. Grid points werealso clustered near the surface of each configuration inorder to adequately resolve the boundary-layer flow. Theamount of grid spacing needed is judged by examiningthe grid spacing parameter,y+, which is given by

(1)

whereρc, uc, andµc are the density, velocity, and viscos-ity at the first cell center next to the solid surface and∆ζis the distance from the first cell center to the body sur-face. Previous research has shown thaty+ values on theorder of 1 provide accurate solutions (ref. 21).

The CFD solutions were obtained using the GeneralAerodynamic Simulation Program (GASP), version 2.2(refs. 22 and 23). GASP is a finite volume code capableof solving the full Reynolds-averaged Navier-Stokes(RANS) equations as well as subsets of these equations,including the parabolized Navier-Stokes (PNS), thin-layer Navier-Stokes (TLNS), and Euler equations. Timeintegration in GASP is based on the integration of primi-tive variables, and convergence to a steady-state solutionis obtained by iterating in pseudotime until the L2 normof the residual vector has been reduced by a sufficientamount. GASP also contains several flux-split algo-rithms and limiters to accelerate convergence to steadystate. Mesh sequencing is available as a means to accel-erate convergence.

In this study, each configuration was modeled as atwo-zone problem, as illustrated in figure 13. The firstzone includes the blunt nose of the configuration. Theflow in this region is a combination of subsonic andsupersonic flow because a small area of subsonic flowexists behind the detached bow shock. Therefore, theTLNS equations are solved over the first zone using aglobal iteration procedure. The second zone encom-passes the remainder of the configuration, extendingfrom the zonal boundary to the base of the configuration.The flow in this region is computed by solving the PNSequations. These equations are valid for regions ofpredominately supersonic flow with no streamwiseseparation. A no-slip boundary condition is applied to allsolid boundaries with a fixed wall temperature of 585°R,which is identical to that specified in the MAXWARPoptimization routine when designing the waveridershapes. Free-stream conditions are applied at the outerboundary, second-order extrapolation from interior cells

y+ρcuc∆ζ

µc------------------=

Page 10: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

8

is applied at the last streamwise plane, and symmetryboundary conditions are applied at the center plane. TheBaldwin-Lomax algebraic turbulence model was used inthese solutions to model turbulent boundary layers,andconvergence to a steady state was obtained by reducingthe L2 norm of the residual vector by 5 orders ofmagnitude.

In order to make appropriate comparisons, the condi-tions at which solutions were obtained were chosenbased on conditions at which experimental data wereavailable. Solutions were obtained at Mach 4.0 at anglesof attack of−6°, 0°, 2°, 4°, and 8° for the straight-wingmodel. Solutions were obtained at Mach 4.0 at angles ofattack of−6°, 0°, and 8° for the cranked-wing model.Solutions were also obtained at off-design Mach num-bers of 2.3 and 4.63 at 0° angle of attack for eachconfiguration.

6. Flow-Field and AerodynamicCharacteristics of Pure Waverider Models

6.1. Flow-Field Characteristics

The flow-field characteristics of the pure waveridermodels at the design Mach number can be illustrated byexamining computational solutions of each configurationand laser vapor-screen photographs from wind-tunneltests. Figure 14 shows a laser vapor-screen photograph ofthe flow at the base of the pure straight-wing waveridermodel and nondimensional static pressure contours at thebase of the same configuration from a CFD solution atMach 4.0, 0° angle of attack, and free-stream Reynoldsnumber of 2.0× 106 per foot. The model lower surface ishighlighted in the photograph by the laser light sheet onthe surface. The bow shock is indicated by the contrastbetween light and dark regions below the light sheet. Onthe left-hand side of the photograph, the shock isobserved to be very near the edge of the lower surface.Thus, the vapor-screen photograph confirms the qualita-tive shock location predicted by the CFD solution. Asmall detachment distance exists even at the design pointcaused by blunt leading edge and boundary-layerdisplacement effects. These effects are not accounted forin the design code. The CFD predictions also indicatethat the high-pressure region remains mostly confinedbelow the model lower surface. A large low-pressureregion (P/P∞ of 0.95 or less) exists near the centerline ofthe model below the bottom surface because of the bot-tom surface expansion present on the model. However,the remainder of the bottom surface flow field is asmooth, conical flow field, so the presence of this slightexpansion surface does not degrade the favorable PAIcharacteristics offered by the waverider. Engine moduleswould be placed upstream of the point where the expan-

sion surface begins, so the flow entering the inlet wouldbe highly compressed. Similar data are shown in fig-ure15 for the cranked-wing pure waverider model. Theshock can be seen in the right-hand side of the photo-graph to be very near the outer leading edge of themodel. The lower surface is again highlighted by thelaser light. The full cross-sectional view is not shownbecause of the poor quality of the photographs. Theexperimental data confirm the qualitative shock locationat the outer leading edge, which is predicted by the CFDsolution for this case as well. Figure 16 further illustratesthat the shock is slightly detached at the outer leadingedge for both models. This figure shows a close-up viewof the outer leading edge at the base of the cranked-wingand straight-wing waverider shapes from CFD solutionsat Mach 4.0 and 0° angle of attack. Both of the viewsin figure 16 are to the same length scale, and non-dimensional static pressure contours are shown in eachview.

The flow-field characteristics of each pure waveridershape at off-design Mach numbers can also be illustratedby examining experimental flow-visualization data andCFD solutions. Figure 17 shows a comparison of avapor-screen photograph and a CFD solution for thecranked-wing shape at Mach 2.3 and 0° angle of attack.The free-stream Reynolds number is 2.0× 106 per foot.The data shown in this figure and orientation of thecamera in the test section are the same as in figures 14and15. At Mach numbers below the design Mach num-ber of 4.0, the shock-wave angle is larger and the detach-ment distance should be much larger than at the designMach number. This outcome is predicted by the CFDsolution and confirmed by the experimental data. Fig-ure18 shows similar views of the same configuration atMach 4.63. The photograph in this figure was taken withthe laser light sheet approximately 5 in. upstream of thebase because the quality of the photograph taken with thelight sheet at the base was poor. At Mach numbersgreater than the design Mach number, the shock movescloser to the leading edge than at the design condition, asillustrated in both the vapor-screen photograph and pre-dicted by the CFD solution. A large high-pressure regionstill exists in the bottom-surface flow field of this config-uration at Mach 4.63. The qualitative shock locations canbe further illustrated by examining planform schlierenphotographs of the cranked-wing model. Figure 19shows schlieren photographs of the cranked-wing purewaverider model in a planform view at Mach 2.3 (top),Mach 4.0 (middle), and Mach 4.63 (bottom). The rightside of the figure shows a close-up view near the leadingedge at each Mach number. The schlieren images in thisfigure have been enhanced by computer imaging tech-niques in order to show the shock structure more clearly.At Mach 2.3, the schlieren photograph shows that the

Page 11: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

9

shock is detached from the leading edge. The outermostshock in the top view represents the bow shock. AtMach4.0, the shock is much closer to the outer leadingedge, but a small detachment distance still exists. AtMach 4.63, the photograph does not show the presence ofa shock wave near the leading edge, possibly because theshock is attached at this condition.

6.2. Aerodynamic Performance

The aerodynamic performance characteristics of thetwo pure waverider models are examined here usingexperimental force and moment data and computationalpredictions. Off-design Mach number and Reynoldsnumber effects are evaluated using experimental data.The longitudinal and lateral-directional stability charac-teristics are also examined for each configuration usingexperimental data. Unless otherwise indicated, all of theexperimental and computational data presented havebeen corrected to a condition of free-stream pressure act-ing at the blunt bases of the configurations, as previouslydiscussed.

The aerodynamic performance of the straight-wingand cranked-wing pure waverider shapes at the designMach number is shown in figures 20 and 21. These fig-ures show experimental data, CFD predictions, anddesign-code predictions for the lift, drag, and lift-dragratios of each configuration at Mach 4.0 and a Reynoldsnumber of 2.0× 106 per foot. The computational valueswere obtained by integrating surface pressure and skinfriction predictions from CFD solutions. Because thedata are corrected to eliminate the base drag, these datashould be interpreted as the performance of the forwardportion (or forebody) of a possible hypersonic configura-tion and not that of a realistic hypersonic vehicle. In gen-eral, agreement is good between the experimental dataand computational predictions. Both the computationalpredictions and experimental data show lower lift andhigher drag values than the predicted design-code values,and these differences can be attributed to several causes.The flow-visualization data and CFD flow-field solutionsshowed that a slight detachment distance exists at theouter leading edge even at the design condition, whichresults in a lift loss and a drag decrease. However, thedesign code assumes an infinitely sharp leading edgewith a perfectly attached shock wave. An additional liftloss results from the expansion ramp on the bottom sur-face of the waverider, and an increase in drag resultsfrom the additional volume added to the upper surface ofthe model. The experimental data also show that themaximum lift-drag ratio occurs near 2° angle of attackfor each configuration. This finding is also consistentwith previous studies, such as those in references 13and16, which show that the maximum lift-drag ratio

occurs at an angle of attack greater than 0° for the wave-rider configurations studied in these references.

A direct comparison of the experimental aero-dynamic performance of the two pure waverider modelsis shown in figure 22. The experimental data show thatthe cranked-wing shape has a slightly higher maximumlift-drag ratio than the straight-wing shape. At positiveangles of attack, the straight-wing shape producedslightly higher lift coefficients. Aside from these obser-vations, there are no significant differences between thetwo configurations.

The off-design performances of the straight-wingpure and cranked-wing pure waverider models are shownin figures 23 and 24, respectively. Each of these figuresshows the experimental lift, drag, and lift-drag ratio at allMach numbers studied as well as maximum lift-dragratio versus Mach number. The data indicate that there isno significant performance degradation at off-designMach numbers. Both configurations show higher maxi-mum lift-drag ratios than the design point value at Machnumbers less than 4.0, using the assumption of free-stream pressure acting at the base. Similar results havebeen found in previous waverider studies (refs.13,16, and 24) and are also typical for non-waveridersupersonic/hypersonic configurations. The cranked-wing waverider shape provides better aerodynamic per-formance at Mach numbers of 4.0 and below. At higherMach numbers, there are no significant differencesbetween the performance of the two configurations.

The effects of Reynolds number on aerodynamicperformance of the straight-wing and cranked-wing con-figurations are shown in figures 25 and 26, respectively.No significant effects of Reynolds number variation wereobserved for either configuration in the range studied,except for a slight increase in maximum lift-drag ratio atthe 3.0× 106-per-foot condition for both configurations.This result is most likely because the skin friction coeffi-cient decreases as Reynolds number increases, resultingin decreased drag and thus increased lift-drag ratios athigher Reynolds numbers. The decrease in drag observedexperimentally is approximately equal to the decrease inviscous drag predicted by the reference temperaturemethod (ref. 11). Computational solutions at Mach 4.0and a Reynolds number of 2.0× 106 per footshow thatthe viscous drag contribution is approximately 34 percentof total drag. By comparison, theMAXWARP designcode predicts a viscous contribution of approximately38 percent to the total drag.

The pitching-moment characteristics of the straight-wing and cranked-wing pure waverider configurationsare shown in figure 27. This figure shows the pitching-moment coefficient versus angle of attack at each Machnumber studied. Both configurations are longitudinally

Page 12: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

10

unstable at all Mach numbers studied. The moment refer-ence center location here is an arbitrarily selectedlocation at the approximate location of the center of grav-ity of the fully integrated model. This moment referencecenter location (16.623 in. aft of the nose) is used for allconfigurations studied unless otherwise stated. Thecranked-wing pitching moment curve is more nonlinearthan that for the straight-wing shape, indicating that theshock may be detached at higher angles of attack for thecranked-wing configuration. The yawing moment char-acteristics are shown in figure 28. This figure shows theyawing moment derivative versus angle of attack at eachMach number studied for both configurations. Thestraight-wing configuration is directionally unstable at allMach numbers studied at angles of attack of 8° andbelow. The cranked-wing configuration is directionallystable at all Mach numbers studied above an angle ofattack of 4°. Both configurations experience a destabiliz-ing effect as Mach number increases. The cranked-wingconfiguration was expected to provide improved direc-tional stability from the increased dihedral along the out-board leading edge. The rolling moment characteristicsare shown in figure 29 for each configuration. Thecranked-wing waverider shows better lateral stabilitycharacteristics than the straight-wing model. Thecranked-wing configuration exhibits positive effectivedihedral above 0° angle of attack at all Mach numbers.The straight-wing model is unstable at angles of attackbelow 6.0° at Mach numbers of 4.0 and 4.63 and is unsta-ble at angles of attack below 4° at a Mach number of 2.3.

7. Component Build-Up Effects

7.1. Effect of Canopy

The effects of adding the canopy on the aerodynamicperformance of the pure straight-wing and cranked-wingwaverider models are illustrated in figures 30 and 31,respectively. These data were obtained for configurationsthat have no control surfaces or engine componentsattached, and the data are corrected to assume free-stream static pressure acting at the base. Each figureshows the lift and drag coefficients as well as lift-dragratios at Mach 4.0 and the maximum lift-drag ratio ateach comparative Mach number studied for the canopy-off and faceted-canopy configurations. The canopy-offconfigurations have the ogive-cylindrical fairing on theupper surface, as discussed previously. Both the straight-wing and cranked-wing configurations show little differ-ence in lift when the canopy is added. The canopy-onconfigurations show slightly higher drag than those withno canopy and an accompanying decrease in lift-dragratios at positive values of lift over the Mach numberrange studied. The maximum lift-drag ratio at Mach 4.0is reduced by 3.6 percent for the straight-wing configura-

tion when the faceted canopy is used. Similarly, a5.1-percent reduction in maximum lift-drag ratio occursfor the cranked-wing configuration. The data indicatethat a penalty was incurred for the canopy, and thereforeattention should be paid to the canopy design in a hyper-sonic waverider-based vehicle.

7.2. Effect of Engine Package

The engine component effects are evaluated by com-paring experimental data from engine-on and engine-offconfigurations. Figures 32 and 33 show the effects ofadding the engine package (ramp, inlet, and nozzle com-ponents) to the straight-wing and cranked-wing configu-rations, respectively. The data shown here are forconfigurations with the canopy and no control surfaces.The data are corrected to assume free-stream static pres-sure acting at the base. No correction is applied to thesedata for the nozzle surface pressures. Each figure showslift and drag coefficients as well as lift-drag ratios atMach 4.0 and the maximum lift-drag ratio at comparativeMach numbers for engine-on and engine-off configura-tions. The addition of engine components results in aslight increase in lift and a significant increase in drag atMach 4.0. These effects are caused by the inlet compres-sion surface and the increase in projected frontal area andproduce a decrease in lift-drag ratio at positive values oflift and a reduction in maximum lift-drag ratio over theMach number range studied. The straight-wing engine-on configuration shows a 19.7-percent reduction in themaximum lift-drag ratio at Mach 4.0 over the engine-off configuration. The cranked-wing model shows a17.7-percent reduction at the same condition.

7.3. Effect of Control Surface Addition

The effects of adding undeflected control surfacesare illustrated by comparing data for configurations withno control surfaces to those with undeflected ailerons andelevons attached. Each configuration includes the canopyand engine components. Data for both the straight-wingand cranked-wing configurations are shown. The coeffi-cient data are reduced by the planform areas of each cor-responding configuration so the effects of increasedplanform are accounted for in the normalization of thesedata. The plots showing drag and lift-drag data includethree separate data sets. The first is the data for thecontrols-off configuration corrected to assume free-stream pressure at the base. Therefore, only forebodydrag values are included in these data and base drag isnot included. The second data set is the controls-on dataand the third set is the controls-off data with base dragincluded (i.e., uncorrected data from wind-tunnel mea-surements), so that these data include the effect of theblunt base. A comparison between the second and thirddata sets shows the aerodynamic effect of adding control

Page 13: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

11

surfaces to the configuration, and a comparison betweenthe first two data sets shows the relative performancebetween the closed configurations and that of the fore-body surface only without the effect of the blunt base.

The effect of adding undeflected control surfaces tostraight-wing waverider configuration with the canopyand engine components attached is summarized in fig-ure 34. The addition of control surfaces causes a slightdecrease in lift coefficient at Mach 4.0. This decrease ispartially caused by the large expansion angle that ispresent on the elevon lower surfaces and a 16-percentincrease in reference area for the controls-on configura-tion. A comparison of the controls-off data with basedrag and the controls-on data shows a decrease in drag ata given lift-coefficient value. There is a slight increase inlift-drag ratios at low positive angles of attack and anincrease in maximum lift-drag ratios when 0° controlsurfaces were added to the configuration. However, acomparison of the controls-on data with the controls-offdata with no base drag shows that the closed configura-tion has significantly higher drag values and lower maxi-mum lift-drag ratios than the forebody-only values. Thisresult indicates that the inclusion of aftbody closure pre-sents a significant challenge in the integration of purewaverider shapes into hypersonic vehicles and that thisaspect of the configuration deserves special consider-ation in the design process. It is likely that the lift-dragratios of a closed configuration cannot approach those ofpure waverider shapes because the effect of base drag isoften not included in lift-drag values for these configura-tions. The effects of control surface addition are similarfor the cranked-wing configuration as indicated in fig-ure35. For reference, the base area is approximately8.3 percent of the planform area for the straight-wingmodel with no control surfaces and approximately9.1 percent for the cranked-wing model.

The control surface design for the configuration usedin this study was a somewhat arbitrary design based onlyon trends from various supersonic fighter aircraft. Amore optimum design could minimize the performancedegradation caused by the closure of the blunt base. Aperformance improvement could be obtained by includ-ing the aftbody closure in the design/optimization pro-cess. Previous studies have examined the possibility ofusing blunt trailing edges on control surfaces as a meansof enhancing the aerodynamic performance (refs. 25to 27). The blunt base reduces the strength of the baserecompression shock and proper design of the trailingedge can result in an increase in base pressure and adecrease in drag. A control surface design that takesadvantage of these effects would enhance the aero-dynamic performance of the configuration. Thisenhancement could be accomplished by reducing thethickness of the base by maintaining the lower surface as

a waverider stream surface all the way to the base whiledesigning the upper surface as an expansion surface.Longer control surfaces would also reduce the closureangle and enhance the pitch control power of theconfiguration.

8. Characteristics of Fully-IntegratedWaverider-Derived Hypersonic CruiseConfigurations

8.1. Aerodynamic Performance

Aerodynamic characteristics of each of the fullyintegrated waverider-derived configurations are exam-ined over the Mach number range using experimentaldata, and the performance of these configurations arecompared to that of the pure waverider shapes. Thefully integrated configurations are defined here to havethe canopy, the engine components, the undeflected aile-rons, the undeflected elevons, and the vertical tailattached. The aerodynamic characteristics of the straight-wing and cranked-wing configurations are presented firstfollowed by comparisons to the corresponding purewaverider configuration.

The aerodynamic performance of the straight-wingand cranked-wing waverider-derived hypersonic cruiseconfigurations are shown in figures 36 and 37, respec-tively. The data presented here have the nozzle surfacepressures corrected to assume free-stream pressure actingon the nozzle surface. The data are presented using thismethod to show the aerodynamic characteristics withoutany propulsive effect on the nozzle surface. The forcedata were corrected by assigning integration areas toeach pressure tap measurement and computing the cor-rected coefficients. The locations of pressure taps areshown in figure 10. The straight-wing configuration hasa maximum lift-drag ratio of 4.69 at Mach 4.0 and thecranked-wing configuration has a value of 4.56 when thenozzle surface pressures are corrected to free-streampressure. The aerodynamic performance of each configu-ration does not vary significantly at off-design Machnumbers. The maximum lift-drag ratio for each configu-ration also occurs near 2° angle of attack at Mach 4.0.The angle of attack for maximum lift-drag ratio increasesas Mach number decreases. At Mach numbers of 2.0 andbelow, the maximum lift-drag ratios for the cranked-wing configuration do not follow the general trend ofincreasing maximum lift-drag ratio with decreasingMach number. This situation results from lift curve slopevalues that show similar inconsistencies at Mach num-bers less than 2.3.

A direct comparison of the straight-wing andcranked-wing fully integrated vehicles is shown in fig-ure 38. The straight-wing configuration produces slightly

Page 14: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

12

higher values of maximum lift-drag ratio than thecranked-wing configuration at Mach numbers of 2.3 andhigher. The straight-wing model also shows higher liftcoefficient values at Mach 4.0. The straight-wing modelshows a maximum lift-drag ratio that is 3.0 percenthigher than that of the cranked-wing configuration at thedesign Mach number of 4.0.

Comparisons of the aerodynamics of the straight-wing pure waverider model and the fully integrated con-figuration are shown in figure 39. This figure shows liftand drag coefficients as well as lift-drag ratios at Mach4.0 and the maximum lift-drag ratios at each Mach num-ber studied. As in figures 34 and 35, these data sets arepresented for comparison in the drag and lift-drag plots.The first data set represents the pure waverider shapewith no base drag included. The second represents thewaverider shape with base drag, and the third representsthe fully integrated configuration. The nozzle surfacepressures are corrected to assume free-stream pressure onthe nozzle surface for the fully integrated vehicles. Acomparison of the pure waverider data with base dragand the fully integrated data shows that the aerodynamicperformance of the pure waverider shape is degradedwhen all of the various vehicle components are added. Areduction in lift coefficient for the fully integrated con-figuration is observed at Mach 4.0 above 0° angle ofattack, which increases as angle of attack increases. Anincrease in drag is observed when all components areintegrated with the pure waverider model. These effectsresult in a decrease in lift-drag ratios at Mach 4.0 and inmaximum lift-drag ratios at comparative Mach numbersof 4.0 and above. At Mach 2.3, there is a slight increasein maximum lift-drag ratio when all vehicle componentsare added. This increase is most likely caused by the noz-zle surface pressure correction to free-stream pressure.The free-stream static pressure increases as Mach num-ber decreases. However, the aerodynamic performanceof the fully integrated vehicle is significantly degradedfrom that of the pure waverider shape only with no basedrag included, because of the drag produced by the con-trol surface addition. The maximum lift-drag ratio atMach 4.0 for the fully integrated vehicle is 4.69, com-pared to 6.68 for the pure waverider shape.

A comparison of the fully integrated cranked-wingconfiguration and the pure cranked-wing waveridermodel yields conclusions similar to those of the compari-son of the straight-wing configurations. Figure 40 showsthe aerodynamic performance of the cranked-wingwaverider forebody and the cranked-wing fully inte-grated configuration. The addition of vehicle compo-nents causes a slight degradation in the aerodynamicperformance, but the lift-drag ratios observed for thefully integrated model are significantly lower than thosefor the pure waverider shape only with no base drag. The

maximum lift-drag ratio at Mach 4.0 for the fully inte-grated configuration is 4.56, compared to a value of 6.72for the fully integrated vehicle.

From these results, it can be concluded that the max-imum lift-drag ratios of a fully integrated waverider-derived configuration with aftbody closure likely cannotapproach those of pure waverider shapes. Theoreticalpredictions for waverider configurations do not includethe effects of aftbody closure. However, it will be shownthat the fully integrated waverider-derived configurationsstudied here are comparable in aerodynamic performanceto previously tested hypersonic models with performanceimprovements possible through enhanced control surfaceand propulsion system designs.

In order to characterize the lift-drag values of theconfigurations studied here, a comparison is madebetween data for the present cranked-wing fully inte-grated waverider-derived configuration and experimentaldata from six hypersonic vehicle wind-tunnel modelspreviously tested in NASA Langley facilities (refs. 28 to33) in figure 41. Although direct comparisons of thesedata are not possible here because of different conditions,geometries, levels of volumetric efficiencies, and forceaccounting methodologies, a range of values can beobtained to compare with the data from the current study.As shown in figure 41, the waverider falls within thesame general range of lift-drag values as the non-waverider hypersonic configurations. The lift-drag ratiosof the waverider configurations studied could beimproved significantly through a better design of the pro-pulsion system and control surface closure. Therefore,the waverider configurations studied here offer at leastcomparable aerodynamic performance and perhaps amodest advantage over conventional non-waveriderhypersonic vehicles.

8.2. Longitudinal Control Effectiveness and Trim

Both fully integrated configurations are longitudi-nally unstable at each Mach number studied. Thepitching-moment coefficient data as a function of angleof attack at each Mach number studied are shown in fig-ure42. Data for the straight-wing and cranked-wing fullyintegrated waverider-derived hypersonic cruise configu-rations are shown. The moment reference center islocated at 62.5 percent of the centerline chord. At higherangles of attack, the cranked-wing configuration shows adestabilizing increase in the pitching moment curve. Thisincrease indicates that the shock may have detached fromthe leading edge of the outer cranked portion of the wingat higher angles of attack. The longitudinal instability ofthese configurations may be addressed in one of twoways. First, it may be possible to shift the center-of-gravity location for a fully integrated flight vehicle to a

Page 15: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

13

location that would provide at least neutral stability overthe Mach number range. Recommendations for suchlocations are presented later in this section. Second, theaddition of a fully functioning propulsion system wouldenhance the longitudinal instability by increasing the aft-body lower surface pressures.

The pitch control effectiveness of the elevons andelevon/aileron combination for the straight-wing config-uration is shown in figure 43. Data are shown for threetrim settings. The first is one with both the elevons andailerons at 0°, the second with a positive 20° elevondeflection (δE) and a 0° aileron deflection (δA), and thethird with both elevons and ailerons deflected at 20°. Theeffectiveness of the elevon decreases as Mach numberincreases, as evidenced by the smaller increments in liftand pitching-moment coefficients produced by eachdeflection. The ailerons were more effective than theelevons in pitch control because of the shadowing of theelevon behind the thick wing shape and the location ofthe elevon in an expansion flow field. The CFD flowfield solutions showed that the bottom surface flow fieldexpands to pressure below free-stream pressure in theregion where the elevons are placed. Also, the closureangle for the elevon was severe because of the thick baseof the waverider. Each aileron has only 70 percent of theplanform area of the elevon but at higher angles of attackgenerates substantially more pitching moment. Thesecharacteristics may be unacceptable and indicate that thepitch control concept should be redesigned.

The pitch control effectiveness of the elevons for thecranked-wing configuration is shown in figure 44. Eachfigure shows data for 0°, 20°, and−20° elevon deflec-tions with 0° ailerons. No runs were made with both aile-rons and elevons deflected at the same angle because ofthe shape of the trailing edge for the cranked-wing con-figuration. The elevon pitch control power for this con-figuration also decreases as Mach number increases.However, in contrast to the straight-wing pitch controldata, the cranked-wing pitching moment curves are non-linear. This factor makes the elevon pitch control powereven more critical for this configuration than for thestraight-wing vehicle. These data indicate that the nose-down pitch control power of this configuration is not suf-ficient. Either symmetric ailerons must be used to pro-vide additional pitch control power or the elevon areashould be increased.

Because of the combination of unstable pitchingmoment characteristics and low pitch control powerobserved in the experimental data, the configurationsshould be balanced such that they are at least neutrallystable to ensure adequate pitch control power throughoutthe angle-of-attack range. For a realistic full-scale flightvehicle, it should be possible to control the center-of-gravity location through packaging. Also, it may be pos-

sible to control the shift in static margin from subsonic tosupersonic speeds using fuel transfer. Neutral stabilitycan be achieved by placing the center of gravity at a loca-tion equal to 58 percent of the centerline chord for thefully integrated straight-wing configuration and 59 per-cent of the centerline chord for the cranked-wing config-uration. Data for lift and pitching-moment coefficientsreferenced to these center-of-gravity locations are shownin figure 45 for the straight-wing vehicle and in figure 46for the cranked-wing vehicle. In figure 45, the data forthe trailing-edge-up elevon deflections were extrapolatedfrom the cranked-wing data and applied to the straight-wing configuration. Also note that all of the data pre-sented here are for unpowered conditions. The additionof a functioning propulsion system will enhance the lon-gitudinal stability of the vehicle even further. These dataare presented only to indicate the effects of an alternativechoice of center-of-gravity locations. Subsequent dataare presented at the original moment reference centerlocation of 62.5 percent of the centerline chord.

8.3. Lateral-Directional Stability and ControlEffectiveness

The lateral-directional stability of the straight-wingand cranked-wing hypersonic cruise vehicles are shownin figures 47 and 48, respectively. Each figure showsyawing and rolling moment derivatives at each Machnumber studied. Both configurations are directionallystable at all Mach numbers investigated, with thecranked-wing model providing higher stability levelsthan the straight-wing model. The cranked-wing fullyintegrated configuration is laterally stable across theangle-of-attack range at all Mach numbers studied. Thestraight-wing fully integrated configuration is laterallyunstable at angles of attack below 6° (at Mach 4.0). Thisroll instability may be caused in part by the high place-ment of the balance in the model. No transfer distance inthe vertical direction was applied to the moment refer-ence center in the presentation of data.

Figures 49 and 50 show the effect of the vertical tailon yawing moment derivative and rolling moment deriv-ative values for each configuration. The effect of the ver-tical tail is to significantly enhance the directionalstability of both the straight-wing and cranked-wing con-figurations, indicated by the large positive shift in yaw-ing moment derivatives when the vertical tail is added toeach model. No rudder control effectiveness runs weremade in this study, so it is not clear whether sufficientyaw control power exists to augment stability. The addi-tion of the vertical tail does not cause any significantchange in the lateral stability characteristics of eitherconfiguration.

The effectiveness of a 20° aileron deflection on thestraight-wing configuration is shown in figure 51. A 20°

Page 16: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

14

aileron deflection indicated here implies one aileron witha 20° trailing-edge-down deflection and the other with a20° trailing-edge-up deflection. The elevons remainedfixed at 0° for these runs. Figure 51 shows rollingmoment and yawing moment increments between thedeflected and nondeflected runs. Additionally, theDATCOM computer code was used to estimate thesteady state roll rates for this configuration (ref. 34).Table 4 shows the steady roll rate capabilities as pre-dicted by this method. The roll rate is shown as deg/secof roll, normalized by flight velocity. For most vehiclesof this type, excess roll-control power is available atlower angles of attack. The requirements for pitch androll control surfaces for the waverider-derived vehiclesmay be driven by low-speed flying qualities. These qual-ities include roll-rate capabilities at subsonic speeds andcrosswind landing requirements.

Figure 52 shows the effectiveness of the ailerons forthe cranked-wing fully integrated configuration. How-ever, a significant difference exists between these resultsand those for the straight-wing configuration. Thecranked-wing ailerons produce considerably moreadverse yaw at 0° angle of attack than the straight-wingconfiguration, as evidenced by the large negative valuesof ∆Cn. The adverse yaw produced by the cranked-wingailerons will further drive the control power requirementsof the rudder.

Figure 53 shows the aileron effectiveness on lateral-directional stability with the ailerons deflected at 20° forthe cranked-wing fully integrated configuration. Rollingmoment and yawing moment increments for a positive20° elevon deflection and a±20° aileron deflection areshown. A comparison of these data shows that a 20°elevon deflection has no effect on roll control power,indicating that interaction between controls is minimal atthe Mach numbers studied here.

9. Concluding Remarks

The aerodynamic performance and stability and con-trol characteristics of two Mach 4.0 waverider-derivedhypersonic cruise configurations were examined. Experi-mental force, moment, and flow-visualization data wereobtained for the two Mach 4.0 waverider configurationsin both test sections of the Langley Unitary Plan WindTunnel (UPWT). The wind-tunnel models were designedto allow testing of various configurations ranging frompure waveriders to fully integrated vehicles. The twopure waverider shapes were referred to as the straight-wing pure and the cranked-wing pure waveriders. Exper-imental data as well as limited computational solutionswere used to examine the flow field and aerodynamiccharacteristics of the two pure waverider shapes, thecomponent build-up effects, and the aerodynamic and

controllability characteristics of the fully integratedhypersonic cruise vehicles.

The flow-field characteristics and aerodynamic per-formance of the two pure waverider shapes wereexamined using experimental and computational data.Computational fluid dynamics (CFD) predictions andlaser vapor-screen photographs of the straight-wing andcranked-wing pure waverider models confirmed theshock attachment/detachment characteristics of eachconfiguration. The shock was slightly detached from theouter leading edge at the design Mach number of 4.0and 0° angle of attack. This detachment distance existsbecause of boundary-layer displacement effects as wellas blunt leading-edge effects. The design code assumesan infinitely sharp leading edge and does not account forthe physical presence of a boundary layer. Comparisonsbetween experimental force data and CFD predictionswere generally good. The maximum lift-drag ratiosobserved experimentally were lower than the design-code predictions, as expected. These lower lift-dragratios were caused by a loss of lift and an increase in dragcaused by the shock not being perfectly attached as wellas to loss of lift from the lower-surface expansion and anincrease in drag from the additional volume added to theupper surface to accommodate model support hardware.The maximum lift-drag ratio for each configurationoccurs at an angle of attack above 0°. Both the CFD pre-dictions and experimental data showed that there were nosignificant performance degradations at off-design Machnumbers. The cranked-wing pure waverider modelexhibited slightly better aerodynamic performance at thecomparative Mach numbers studied than the straight-wing model.

Component build-up effects of waverider-derivedvehicles were examined by comparing experimentalforce and moment data. The primary effect of individu-ally adding the canopy and the engine package was toincrease the drag of the configuration, thereby resultingin a degradation in aerodynamic performance. The aero-dynamic effect of adding control surfaces was to increasethe maximum lift-drag ratios slightly at each Machnumber studied. However, the aerodynamic performanceof the controls-on configurations was significantlydegraded from that of the pure waverider shape only bythe addition of aftbody closure and the associated dragproduction. The values presented for the pure waveridermodel show the performance of the waverider surfaceonly and do not include base drag. These results indicatethat additional consideration should be applied tothe design of control surfaces and aftbody closure inwaverider-based hypersonic cruise configurations. Acontrol surface configuration with a less severe closureangle or controls with blunt trailing edges may result inimproved performance. Inclusion of the aftbody closure

Page 17: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

15

in the optimization process for the waverider shape mayalso improve the performance significantly.

The characteristics of the fully integrated waverider-derived hypersonic cruise vehicles were also examinedby comparisons of experimental force and moment data.The aerodynamic performance of each fully integratedwaverider model (straight-wing and cranked-wing con-figuration) was significantly degraded from that of thepure waverider shapes, because of the inclusion of aft-body closure in the fully integrated configuration. Thestraight-wing fully integrated configuration providedslightly better aerodynamic performance than thecranked-wing fully integrated model. The maximum lift-drag ratios at Mach 4.0 were 4.69 for the straight-wingmodel and 4.56 for the cranked-wing model. The wave-rider concept also provides a uniform compressed flowfield to the inlet, which offers potential advantages forairbreathing propulsion systems integration. The use ofdifferent generating flow fields, such as osculating-coneand cone-wedge flow fields, may further improve thesecharacteristics. Furthermore, the results of this studyhave identified areas where design improvements couldenhance performance, such as control surfaces, aftbodyclosure, and propulsion system design.

Both fully integrated vehicles are longitudinallyunstable across the Mach number range studied forunpowered conditions with the selected referencemoment center. Additionally, locations were recom-mended for placement of the center of gravity in eachconfiguration in order to ensure at least neutral stabilityacross the Mach number range. The pitch-control effec-tiveness of the elevons was judged to be unacceptable forboth configurations, and the data indicate that the pitchcontrol concept should be redesigned. The ailerons weresignificantly more effective than the elevons for pitchcontrol. The cranked-wing vehicle shows significantlybetter lateral-directional stability than the straight-wingvehicle. The straight-wing configuration was unstable atangles of attack below 6° at Mach 4.0. The vertical tailhas a significant stabilizing effect on directional stability,but very little effect on lateral stability. The ailerons arealso highly effective for the cranked-wing vehicle, butproduce a significant amount of adverse yaw.

NASA Langley Research CenterHampton, VA 23681-0001May 6, 1996

References

1. Bowcutt, K. G.; Anderson, J. D.; and Capriotti, D.: ViscousOptimized Hypersonic Waveriders. AIAA-87-0272, Jan. 1987.

2. Corda, Stephen; and Anderson, John D., Jr.: Viscous Opti-mized Hypersonic Waveriders Designed From AxisymmetricFlow Fields. AIAA-88-0369, Jan. 1988.

3. Eggers, A. J., Jr.; Ashley, Holt; Springer, George S.; Bowles,Jeffrey V.; and Ardema, Mark D.: Hypersonic Waverider Con-figurations From the 1950’s to the 1990’s. Proceedings of the1st International Hypersonic Waverider Symposium, John D.Anderson, Jr., Mark J. Lewis, Stephen Corda, and Isaiah M.Blankson, eds., Univ. of Maryland, 1990.

4. Sobieczky, H.; Dougherty, F. C.; and Jones, K.: HypersonicWaverider Design From Given Shock Waves.Proceedings ofthe 1st International Hypersonic Waverider Symposium,JohnD. Anderson, Jr., Mark J. Lewis, Stephen Corda, andIsaiah M. Blankson, eds., Univ. of Maryland, 1990.

5. Takashima, Naruhisa; and Lewis, Mark J.: Waverider Configu-rations Based on Non-Axisymmetric Flow Fields for Engine-Airframe Integration. AIAA-94-0380, Jan. 1994.

6. Rasmussen, Maurice L.: Waverider Configurations DerivedFrom Inclined Circular and Elliptic Cones.J. Spacecr. &Rockets, vol. 17, no. 6, Nov.–Dec. 1980, pp. 537–545.

7. Kuchemann, D.:The Aerodynamic Design of Aircraft. Perga-mon Press, Inc., 1978.

8. O’Neill, Mary K. L.; and Lewis, Mark J.: Optimized ScramjetIntegration on a Waverider.J. Aircr., vol. 29, no. 6, Nov.–Dec.1992, pp. 1114–1121.

9. Corda, Stephen; and Seifert, E. Scott:User Information forMaryland Axisymmetric Wave Rider Program (MAXIWARP).Univ. of Maryland, Jan. 1989.

10. Anderson, John D., Jr.: Modern Compressible Flow—WithHistorical Perspective. McGraw-Hill, Inc., 1982.

11. Anderson, John D., Jr.: Hypersonic and High Temperature GasDynamics.McGraw-Hill, Inc., 1989.

12. Beuerlein, David L.: Optimization of Waveriders to MaximizeMission Performance.Proceedings of the 1st InternationalHypersonic Waverider Symposium, John D. Anderson, Jr.,Mark J. Lewis, Stephen Corda, and Isaiah M. Blankson, eds.,Univ. of Maryland, 1990.

13. Cockrell, Charles E., Jr.: Interpretation of Waverider Perfor-mance Data Using Computational Fluid Dynamics. AIAA-93-2921, July 1993.

14. Jackson, C. M., Jr.; Corlett, W. A.; and Monta, W. J.:Descrip-tion and Calibration of the Langley Unitary Plan Wind Tunnel.NASA TP-1905, 1981.

15. Cockrell, Charles Edward, Jr.: Vehicle Integration Effects onHypersonic Waveriders.NASA TM-109739, 1994.

16. Bauer, Steven X. S.; Covell, Peter F.; Forrest, Dana K.; andMcGrath, Brian E.: Analysis of Two Viscous OptimizedWaveriders.Proceedings of the 1st International HypersonicWaverider Symposium, John D. Anderson, Jr., Mark J. Lewis,Stephen Corda, and Isaiah M. Blankson, eds., Univ. ofMaryland, 1990.

Page 18: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

16

17. Braslow, A. L.; and Knox, Eugene C.:Simplified Method forDetermination of Critical Height of Distributed RoughnessParticles for Boundary-Layer Transition at Mach NumbersFrom 0 to 5.NACA TN-4363, Sept. 1958.

18. Braslow, A. L.; Harris, R. V., Jr.; and Hicks, R. M.:Use ofGrit-Type Boundary-Layer-Transition Trips on Wind-TunnelModels.NASA TN D-3579, 1966.

19. Stallings, Robert L., Jr.; and Lamb, Milton:Effects of Rough-ness Size on the Position of Boundary-Layer Transition and onthe Aerodynamic Characteristics of a 55 Degree Swept DeltaWing at Supersonic Speeds.NASA TP-1027, 1977.

20. Steinbrenner, John P.; and Chawner, John R.:The GRIDGENVersion 8 Multiple Block Grid Generation Software. MDAEngineering Report 92-01, 1992.

21. Richardson, Pamela F.; McClinton, Charles R.; Bittner, RobertD.; Dilley, A. Douglas; and Edwards, Kelvin W.: HypersonicCFD Applications for the National Aero-Space Plane. SAE892310, Sept. 1989.

22. McGrory, W. D.; Huebner, L. D.; Slack, D. C.; and Walters,R. W.: Development and Application of GASP 2.0. AIAA-92-5067, Dec. 1992.

23. McGrory, William D.; Slack, David C.; Applebaum,Michael P.; and Walters, Robert W.: GASP Version 2.2—TheGeneral Aerodynamic Simulation Program. AeroSoft, Inc.,1993.

24. Takashima, Naruhisa:Navier-Stokes Computations of a Vis-cous Optimized Waverider. NASA CR-189658, 1992.

25. Bushnell, D. M.: Supersonic Aircraft Drag Reduction. AIAA-90-1596, June 1990.

26. Chapman, Dean R.:Reduction of Profile Drag at SupersonicVelocities by the Use of Airfoil Sections Having a BluntTrailing Edge. NACA TN-3503, 1955. (Supersedes NACARM A9H11.)

27. Dutton, J.; Herrin, J.; Molezzi, M.; Mathur, T.; and Smith, K.:Some Problems in Transonic Aerodynamics. AIAA-95-0476,Jan. 1995.

28. Penland, Jim A.; Hallissy, James B.; and Dillon, James L.:Aerodynamic Characteristics of a Hypersonic Research Air-plane Concept Having a 70° Swept Double-Delta Wing atMach Numbers From 0.80 to 1.20, With Summary of DataFrom 0.20 to 6.0.NASA TP-1552, 1979.

29. Dillon, James L.; and Pittman, Jimmy L.:Aerodynamic Char-acteristics at Mach 6 of a Wing-Body Concept for a Hyper-sonic Research Airplane.NASA TP-1249, 1978.

30. Penland, Jim A.; Edwards, Clyde L. W.; Witcofski, Robert D.;and Marcum, Don C., Jr.: Comparative Aerodynamic Study ofTwo Hypersonic Cruise Aircraft Configurations Derived FromTrade-Off Studies.NASA TM X-1436, 1967.

31. Small, William J.; Kirkham, Frank S.; and Fetterman,David E.: Aerodynamic Characteristics of a HypersonicTransport Configuration at Mach 6.86. NASA TN-D-5885,1970.

32. Watts, Joe D.; Olinger, Frank V.; Weidner, John P.; Johnson,Stuart K.; Sanders, Bobby W.; and Keyes, J. Wayne: Mach 5Cruise Aircraft Research.Proceedings of the Langley Sympo-sium on Aerodynamics,Volume 2, Sharon H. Stack, Compiler,NASA CP-2398, 1986, pp. 285–304.

33. Ellison, James C.:Investigation of the Aerodynamic Charac-teristics of a Hypersonic Transport Model at Mach Numbersto 6.NASA TN D-6191, 1971.

34. Williams, John E.; and Vukelich, Steven R.: The USAFStability and Control Digital Datcom. Volume 1: User’s Man-ual. AFFDL-TR-79-3032-VOL-1, Apr. 1979. (SupersedesAFFDL-TR-73-23, AFFDL-TR-74-68, and AFFDL-TR-76-45-VOL-1.)

Page 19: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

17

Table 1. Characteristics of Straight-Wing Waverider Designed by MAXWARP

Waverider length, in. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .24.0

Span/length. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.83

Base height/length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.092

Volumetric efficiency (Veff). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.112

Planform area,Sref, ft2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.89

Predicted maximumL/D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.9

Base area, ft2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.136

Table 2. Characteristics of Cranked-Wing Waverider Designed by MAXWARP

Waverider length, in. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .24.0

Span/length. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.96

Base height/length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.092

Volumetric efficiency (Veff). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.108

Planform area,Sref, ft2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.05

Predicted maximumL/D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.7

Base area, ft2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .0.153

aFor some data: 58.0.bFor some data: 59.0.

Table 3. Reference Quantities for Various Configurations

Configuration Sref, ft2 Span, in.

Length, in. Basearea, ft2

Xc.g.,percent of

Straight-wing pure model 1.894 19.80 24.0 0.1580 69.3

Straight-wing pure model with enginecomponents

1.894 19.80 24.0 0.1481 69.3

Straight-wing fully integrated model 2.202 19.80 26.60 0.0194 62.5a

Cranked-wing pure model 2.052 23.016 24.0 0.1860 69.3

Cranked-wing pure model with enginecomponents

2.052 23.016 24.0 0.1745 69.3

Cranked-wing fully integrated model 2.346 23.016 26.60 0.0194 62.5b

Table 4. Steady-Roll-Rate Capabilities Calculated FromDATCOM for Straight-Wing Fully

Integrated Configuration

Mach number Pss per unit velocity,

2.3 0.1194.0 0.0954.63 0.095

c c

deg sec⁄ft sec⁄

--------------------

Page 20: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

18

Figure 1. Design of conical-flow-derived waverider.

Figure 2. Comparison ofL/Dmax values of conventional vehicles and waveriders.

Y

Z

X

Waverider leading edge

Bottom surface (stream surface)

Conical shock wave

M∞

5 10 15 20 25 300

2

4

6

8

10

12

14

16

18

L/D

max

L/Dmax = 4(M + 3)/M "L/D Barrier" Conical-flow-derived waveriders

Mach number

Page 21: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

19

Figure 3. Straight-wing pure waverider shape designed by MAXWARP.

Figure 4. Cranked-wing pure waverider shape designed by MAXWARP.

Planform

Profile

Oblique

Rear

Planform

Profile

Oblique

Rear

Page 22: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

20

Figure 5. Straight-wing pure waverider model in UPWT.

Page 23: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

21

Fig

ure

6. C

rank

ed-w

ing

pure

wav

erid

er m

odel

in U

PW

T.

Page 24: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

22

Figure 7. Lower surface of cranked-wing pure waverider model.

Expansion surface

Page 25: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

23

Fig

ure

8. S

trai

ght-

win

g fu

lly in

tegr

ated

wav

erid

er m

odel

.

Ver

tical

tail

Ele

von

surf

ace

(und

efle

cted

)

Aile

ron

surf

ace

defl

ecte

d at

+20

°

Eng

ine

pack

age

Can

opy

Page 26: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

24

Fig

ure

9. F

ully

inte

grat

ed m

odel

with

var

ious

com

pone

nts.

Cra

nked

w

ingt

ips

with

aile

rons

Smoo

th o

give

fai

ring

Stra

ight

-win

g m

odel

with

0° e

levo

ns

and

+20

° aile

rons

atta

ched

Eng

ine

off-

bloc

k

Can

opy

Ele

vons

Aile

rons

Page 27: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

25

(a) Short expansion ramp used with no-controls configurations.

Figure 10. Three-view drawings of expansion ramps.

Tap No. Model Sta. B.L. W.L.1� 22.030 –1.353 –1.7852� 22.030 –1.015 –1.7883� 22.030 –0.676 –1.7924� 22.030 –0.338 –1.7965� 22.791 –1.353 –1.2446� 22.791 –1.015 –1.2527� 22.791 –0.676 –1.2548� 22.791 –0.338 –1.2559� 23.552 –1.353 –0.964

10� 23.552 –1.015 –0.97411� 23.552 –0.676 –0.97412� 23.552 –0.338 –0.975

3.5"

Pressure tap locations on nozzle surface

Rear view

Top view

B.L. 0.0

1.971"

M.S.21.779"

M.S.23.750"

1.386"

W.L. 0.0

Side view

B.L. 0.0o oooo oooo ooo

Taps 1-4

Taps 5-8

Taps 9-12

Short ramp (No-controls configurations)

Page 28: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

26

(b) Long expansion ramp used with fully integrated configurations.

Figure 10. Concluded.

2.011"

4.818"

M.S.21.779

M.S.26.597

W.L. 0.0

3.5"

B.L. 0.0

B.L. 0.0Sting sleeve

Side view Rear view

Top view

o o o oo o o oo o o oo o o oo o oo o

Pressure tap locations on nozzle surface

Taps 1-4

Taps 5-8

Taps 9-12Taps 13-16

Taps 17-19

Taps 21-22

Taps 20, 23, 24 located on sting sleeve

W.L. 0.0

Tap No. Model Sta. B.L. W.L.13� 24.313 –1.353 –0.72414� 24.313 –1.015 –0.73115� 24.313 –0.676 –0.73416� 24.313 –0.338 –0.74517� 25.075 –1.353 –0.51618� 25.075 –1.015 –0.52319� 25.075 –0.676 –0.53220� 25.075 –0.338 –0.73921� 25.836 –1.353 –0.33222� 25.836 –1.015 –0.33923� 25.836 –0.676 –0.45124� 25.836 –0.338 –0.739

Locations for taps 1 to 12 are given on previous page.

Long ramp (Fully integrated configurations)

Page 29: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

27

(a) Elevons. (b) Straight-wing ailerons.

(c) Cranked-wing inboard ailerons. (d) Cranked-wing outboard ailerons.

Figure 11. Dimensions of elevons and ailerons.

4.389"

1.750"10.8°

3.22"

Hinge line

Inboard side

Outboard side

Trailing edge

2.66"

Hinge line 3.760"

Trailing edge

0.588"

10°

2.66"

Hinge line

2.047"

Trailing edge

2.047"

10°

2.615"

Hinge line

3.323"

Trailing edge

2.570"

30°

Page 30: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

28

Figure 12. Three-view drawing of fully integrated waverider model.

26.6"

11.5"

Cranked wing

Straight wing

Cranked wing

Straight wing

Canopy

Waverider stream surface Compression

surface

9.9"Moment ref. center 62.5 percent of model length

Sidewalls Aileron surface

Elevon surface

Vertical tailCylindrical uppper surface fairing

Nozzle surface

Lower surface of engine module

2.81"

0.87" radius

5.029"

Page 31: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

29

Figure 13. Coordinates and computational scheme for waverider CFD solutions.

η ξ

ζ

PNS marching zone

Z

X

Y

TLNS global iteration zone

Page 32: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

30

(a) Vapor-screen photograph of base.

(b) Base view of CFD solution.

Figure 14. Comparison of base-view vapor-screen photograph and CFD nondimensional static pressure contours ofstraight-wing pure waverider model atM = 4.0 andα = 0°.

Shock

Lower surface of model

P/P∞

1.71

1.62

1.54

1.46

1.37

1.29

1.20

1.12

1.03

0.95

Base of configuration

Page 33: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

31

(a) Vapor-screen photograph of base.

(b) Base view of CFD solution.

Figure 15. Comparison of base-view vapor-screen photograph and CFD nondimensional static pressure contours ofpure cranked-wing waverider model atM = 4.0 andα = 0°.

Shock

Lower surface of model

1.54

1.47

1.39

1.32

1.24

1.17

1.10

1.02

0.95

1.61

P/P∞Base of

configuration

Page 34: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

32

(a) Cranked-wing pure waverider model.

(b) Straight-wing pure waverider model.

Figure 16. Comparison of CFD nondimensional static pressure contours near leading edge at base of cranked-wing andstraight-wing pure waverider models atM = 4.0 andα = 0°.

P/P∞

2.18

2.05

1.92

1.79

1.66

1.53

1.39

1.26

1.13

1.00

P/P∞

2.18

2.05

1.92

1.79

1.66

1.53

1.39

1.26

1.13

1.00

Page 35: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

33

(a) Vapor-screen photograph at base.

(b) Base view of CFD solution.

Figure 17. Comparison of base-view vapor-screen photograph and CFD nondimensional static pressure contours ofcranked-wing pure waverider model atM = 2.3 andα = 0°.

Shock

Upper surface of model

P/P∞

1.24

1.21

1.17

1.13

1.10

1.06

1.02

0.99

0.95

1.28Base of configuration

Page 36: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

34

(a) Vapor-screen photograph 5 in. upstream of base.

(b) Base view of CFD solution.

Figure 18. Comparison of base-view vapor-screen photograph and CFD nondimensional static pressure contours ofcranked-wing pure waverider model atM = 4.63 andα = 0°.

Shock

Upper surface of model

P/P∞

1.62

1.54

1.46

1.37

1.29

1.20

1.12

1.03

0.95

1.71Base of configuration

Page 37: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

35

Figure 19. Comparison of planform schlieren photographs of cranked-wing pure waverider model atM = 2.3, 4.0,and 4.63.

Mach 2.30

Mach 4.00

Mach 4.63

Shock wave

Shock wave

Page 38: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

36

Figure 20. Comparison of experimental data, CFD predictions, and design-code predictions for aerodynamic perfor-mance of straight-wing pure waverider model atM = 4.0 and Reynolds number of 2.0× 106 per foot.

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

Experimental

CFD predictions

MAXWARP prediction

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Experimental

CFD predictions

MAXWARP prediction

Experimental

CFD predictions

MAXWARP prediction

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

DL

ift-

drag

rat

io, L

/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

Page 39: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

37

Figure 21. Comparison of experimental data, CFD predictions, and design-code predictions for aerodynamic perfor-mance of cranked-wing pure waverider model atM = 4.0 and Reynolds number of 2.0× 106 per foot.

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

Experimental

CFD predictions

MAXWARP prediction

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Experimental

CFD predictions

MAXWARP prediction

Experimental

CFD predictions

MAXWARP prediction

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

DL

ift-

drag

rat

io, L

/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

Page 40: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

38

Figure 22. Comparison of aerodynamic performance of straight-wing and cranked-wing pure configurations atM = 4.0and Reynolds number of 2.0× 106 per foot.

Straight wing

Cranked wing

Straight wing

Cranked wing

Straight wing

Cranked wing

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

DL

ift-

drag

rat

io, L

/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

Page 41: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

39

Figure 23. Aerodynamic performance of straight-wing pure waverider configuration across Mach number rangestudied.

.4–10

–8

–6

–4

–2

0

2

4

6

8

10

12–.2

–.1

0

.1

.2

.3

.4

Mach 2.3Mach 3.5

Mach 4.63

Mach 4.0

Mach 4.2

α.4

0

.01

.02

.03

.04

.05

.06

.07

5.05.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0

Mach 2.3Mach 3.5

Mach 4.63

Mach 4.0

Mach 4.2

Mach 2.3Mach 3.5

Mach 4.63

Mach 4.0

Mach 4.2

1086420–2–4–6–8 .3.2.10–.1–.2

Lift coefficient, CL

4.54.03.53.02.52.0

Mach number, M∞

.3.2.10–2 –.1

Lif

t-dr

ag r

atio

, L/D

Lif

t coe

ffic

ient

, CL

Max

imum

L/D

Dra

g co

effi

cien

t, C

D

Lift coefficient, CL

Page 42: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

40

Figure 24. Aerodynamic performance of cranked-wing pure waverider configuration across Mach number rangestudied.

Mach 3.5

Mach 4.0

Mach 4.2

Mach 4.63

Mach 2.3

Mach 1.6

Mach 3.5

Mach 4.0

Mach 4.2

Mach 4.63

Mach 2.3

Mach 1.6

Mach 3.5

Mach 4.0

Mach 4.2

Mach 4.63

Mach 2.3

Mach 1.6

.4–14

–8–6–4–2

02468

14

12–.2

–.1

0

.1

.2

.3

.4

α.4

0

.01

.02

.03

.04

.05

.06

.07

5.05.05.5

6.5

7.5

8.5

9.510.010.511.0

1086420–2–4–6–8 .3.2.10–.1–.2

Lift coefficient, CL

4.54.03.53.02.51.5

Mach number, M∞

.3.2.10–2 –.1

Lif

t-dr

ag r

atio

, L/D

Lif

t coe

ffic

ient

, CL

Max

imum

L/D

Dra

g co

effi

cien

t, C

D

9.0

8.0

7.0

6.0

1012

–10–12

2.0

Lift coefficient, CL

11.512.0

Page 43: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

41

Figure 25. Effects of Reynolds number on aerodynamic performance of straight-wing pure waverider configuration atM = 4.0.

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–10

–6

–4

–2

0

2

4

6

10

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

DL

ift-

drag

rat

io, L

/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

8

–8

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

Page 44: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

42

Figure 26. Effects of Reynolds number on aerodynamic performance of cranked-wing pure waverider configuration atM = 4.0.

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–10

–6

–4

–2

0

2

4

6

10

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

DL

ift-

drag

rat

io, L

/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

8

–8

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

1.5 × 106 per foot

3.0 × 106 per foot

2.0 × 106 per foot

Page 45: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

43

(a) Straight wing.

(b) Cranked wing.

Figure 27. Pitching moment characteristics of pure waverider configurations.

Mach 2.3

Mach 3.5

Mach 4.0

Mach 4.2

Mach 4.63

12α

1086420–2–4–6–8

.030

.025

.020

.015

.010

.005

0

–.005

–.010

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Mach 2.3

Mach 3.5

Mach 4.0

Mach 4.2

Mach 4.63

Mach 1.6

12α

1086420–2–4–6–8

.030

.025

.020

.015

.010

.005

0

–.005

–.010

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Page 46: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

44

(a) Straight wing.

(b) Cranked wing.

Figure 28. Yawing moment characteristics of pure waverider configurations.

Stable

Mach 2.3

Mach 4.0

Mach 4.63

12α

1086420–2–4–6

.0004

.0003

.0002

.0001

0

–.0002

–.0003

–.0004

Yaw

ing-

mom

ent d

eriv

ativ

e, C

–.0001

Stable

Mach 2.3

Mach 4.0

Mach 4.63

Mach 1.6

12α

1086420–2–4–6

.0008

.0006

.0004

.0002

0

–.0002

–.0004

–.0006

Yaw

ing-

mom

ent d

eriv

ativ

e, C

Page 47: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

45

(a) Straight wing.

(b) Cranked wing.

Figure 29. Rolling moment characteristics of pure waverider configurations.

Stable

Mach 2.3

Mach 4.0

Mach 4.63

12α

1086420–2–4

.0016

.0012

0

–.0004

–.0008

–.0012

Rol

ling-

mom

ent d

eriv

ativ

e, C

lβ.0008

.0004

Stable

Mach 2.3

Mach 4.0

Mach 4.63

Mach 1.6

12α

1086420–2–4

.0008

.0004

0

–.0020

–.0024

–.0028

Rol

ling-

mom

ent d

eriv

ativ

e, C

–.0016

–.0012

–.0008

–.0004

Page 48: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

46

Figure 30. Effect of canopy on aerodynamic performance of straight-wing pure waverider configuration.

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

Canopy off

Canopy on

Canopy off

Canopy on

Canopy off

Canopy on

Canopy off

Canopy on

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.05.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0

4.54.03.53.02.52.0

Mach number, M∞

Max

imum

L/D

Page 49: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

47

Figure 31. Effect of canopy on aerodynamic performance of cranked-wing pure waverider configuration.

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

Canopy off

Canopy on

Canopy off

Canopy on

Canopy off

Canopy on

Canopy off

Canopy on

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.05

6

7

8

9

10

12

11

12

13

14

4.54.03.53.02.51.5

Mach number, M∞

Max

imum

L/D

2.0

Page 50: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

48

Figure 32. Effect of adding engine package on aerodynamics of straight-wing pure waverider model with canopy.

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

Engine off

Engine on

Engine off

Engine on

Engine off

Engine on

Engine off

Engine on

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.04

6

7

8

9

10

4.54.03.53.02.52.0

Mach number, M∞

Max

imum

L/D

5

Page 51: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

49

Figure 33. Effect of adding engine package on aerodynamics of cranked-wing pure waverider model with canopy.

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

Engine off

Engine on

Engine off

Engine on

Engine off

Engine on

Engine off

Engine on

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.04

6

7

8

9

10

4.54.03.53.02.51.5

Mach number, M∞

Max

imum

L/D

2.0

5

Page 52: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

50

Figure 34. Effect of undeflected control surface addition on aerodynamics of straight-wing pure waverider configura-tion with canopy and engine package attached.

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.04.0

4.5

5.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

9.0

4.54.03.53.02.52.0

Mach number, M∞

Max

imum

L/D

Page 53: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

51

Figure 35. Effect of undeflected control surface addition on aerodynamics of cranked-wing pure waverider configura-tion with canopy and engine package attached.

M∞ = 4.0

M∞ = 4.0M∞ = 4.0

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.0

4.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

4.54.03.53.02.51.5

Mach number, M∞

Max

imum

L/D

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

No controls (no base drag)0° controlsNo controls (base drag)

2.0

5.0

4.5

3.5

Page 54: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

52

Figure 36. Aerodynamics of fully integrated straight-wing configuration.

.30

.03

.04

.05

.06

.07

.08

.09

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.2.10–.2

.3.2.10–.1–.2

1086420–2–4–6–8

5.03.0

5.5

6.0

6.5

7.0

7.5

8.0

3.5

4.0

4.5

5.0

4.54.03.53.02.52.0

Mach number, M∞

Max

imum

L/D

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

.02

.01

–.1

Page 55: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

53

Figure 37. Aerodynamics of fully integrated cranked-wing configuration.

.40

.01

.02

.03

.04

.05

.09

.10

Lift coefficient, CL

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

–6

–4

–2

0

2

4

6

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.3.10–.1–.2–.31086420–2–4–6–8

5.03.0

5.5

6.0

6.5

7.0

7.5

8.0

4.54.03.53.02.51.5Mach number, M∞

Max

imum

L/D

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

5.0

4.5

4.0

3.5

2.0

.2

.06

.07

.08M∞ = 2.0

M∞ = 1.8

M∞ = 1.6

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 2.0

M∞ = 1.8

M∞ = 1.6

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 2.0

M∞ = 1.8

M∞ = 1.6

.4Lift coefficient, CL

.3.10–.1–.2–.3 .2

Page 56: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

54

Figure 38. Comparison of aerodynamics of straight-wing and cranked-wing fully integrated configurations.

Straight wing

Cranked wing

Straight wing

Cranked wing

Straight wing

Cranked wing

Straight wing

Cranked wing

.300

.01

.02

.03

.04

.05

.06

.07

Lift coefficient, CL

12–.20

–.15

–.10

–.05

0

.05

.10

.15

.20

.25

.30

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.25.20.15.10.050–.05–.10

.30.25.20.15.10.050–.05–.10

1086420–2–4–6–8

5.03

4

5

6

7

8

9

10

11

12

4.54.03.53.02.51.5Mach number, M∞

Max

imum

L/D

M∞ = 4.0 M∞ = 4.0

M∞ = 4.0

2.0

Page 57: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

55

Figure 39. Comparison of aerodynamic performance of straight-wing fully integrated and straight-wing pure waveriderconfigurations.

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

.30

.01

.02

.03

.04

.05

.06

.09

Lift coefficient, CL

12–.3

–.1

0

.1

.2

.3

.4

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.2.10–.1–.2

.3.2.10–.1–.2

1086420–2–4–6–8

5.03.0

5.5

6.0

6.5

7.0

7.5

8.08.5

9.0

4.54.03.53.02.52.0Mach number, M∞

Max

imum

L/D

M∞ = 4.0 M∞ = 4.0

M∞ = 4.0

.07

.08

–.2

5.0

4.5

4.0

3.5

Page 58: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

56

Figure 40. Comparison of aerodynamic performance of cranked-wing fully integrated and cranked-wing pure waveriderconfigurations.

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

Pure waverider (base drag)Fully integrated

Pure waverider (no base drag)

.30

.01

.02

.03

.04

.05

.06

.09

Lift coefficient, CL

12–.3

–.1

0

.1

.3

.4

α

–8

–6

–4

–2

0

2

4

6

8

Lift coefficient, CL

Lif

t coe

ffic

ient

, CL

Dra

g co

effi

cien

t, C

D

Lif

t-dr

ag r

atio

, L/D

.2.10–.2

.3.2.10–.1–.2

1086420–2–4–6–8

5.03

4

5

6

7

8

9

10

11

12

4.54.03.53.02.51.5Mach number, M∞

Max

imum

L/D

M∞ = 4.0 M∞ = 4.0

M∞ = 4.0

.07

.08

–.1

.2

–.2

2.0

Page 59: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

57

Figure 41. Assessment of aerodynamic performance of waverider-derived vehicle.

82

3

4

5

6

765431Mach number, M∞

Max

imum

L/D

Present fully integrated cranked-wing waveriderReference 28Reference 29Reference 30Reference 31Reference 32Reference 33

2

Page 60: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

58

(a) Straight wing.

(b) Cranked wing.

Figure 42. Longitudinal stability of each fully integrated waverider-derived configuration with undeflected elevons andailerons.

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

12–.010

–.005

0

.005

.010

.015

.020

α

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

1086420–2–4–6–8

M∞ = 1.6

M∞ = 1.8

M∞ = 2.0

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

12–.005

0

.005

.010

.015

.020

.025

α

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

1086420–2–4–6–8

Page 61: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

59

(a) M∞ = 2.3 and 4.0.

Figure 43. Pitch control effectiveness of elevons and elevon/aileron combination for straight-wing fully integratedconfiguration.

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

–0.15

–.010

–.005

0

.005

.015

.020

.025

12–.2

–.1

0

.1

.2

.3

α

–8

–6

–4

–2

0

2

4

6

8

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.015

–.010

–.005

0

.005

.010

.015

.020

.025

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 2.3 M∞ = 2.3

M∞ = 4.0

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 4.0

.010

Page 62: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

60

(b) M∞ = 4.63.

Figure 43. Concluded.

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 20°, δE = 20°

–0.15

–.010

–.005

0

.005

.015

.020

.025

12–.2

–.1

0

.1

.2

.3

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

1086420–2–4–6–8

M∞ = 4.63 M∞ = 4.63

12α

1086420–2–4–6–8

.010

Page 63: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

61

(a) M∞ = 1.6 and 1.8.

Figure 44. Pitch control effectiveness of elevons for cranked-wing fully integrated configuration.

–.015

–.010

0

.010

.015

.030

.035

.040

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.015

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 1.6 M∞ = 1.6

M∞ = 1.8

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 1.8

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

.025

.020

.005

–.005

.005

–.3

–.2

–.1

.1

.2

.3

.4

0

Page 64: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

62

(b) M∞ = 2.0 and 2.3.

Figure 44. Continued.

–.015

–.010

0

.010

.015

.030

.035

.040

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.015

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 2.0 M∞ = 2.0

M∞ = 2.3

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 2.3

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

.025

.020

.005

–.005

.005

–.3

–.2

–.1

.1

.2

.3

.4

0

Page 65: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

63

(c) M∞ = 4.0 and 4.63.

Figure 44. Concluded.

–.015

–.010

0

.010

.015

.030

.035

.040

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.015

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 4.0 M∞ = 4.0

M∞ = 4.63

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 4.63

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

.025

.020

.005

–.005

.005

–.3

–.2

–.1

.1

.2

.3

.4

0

Page 66: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

64

(a) M∞ = 2.3 and 4.0.

Figure 45. Pitch control effectiveness of elevons for straight-wing fully integrated configuration with moment referencecenter at 58 percent of body length.

–.015

–.010

0

.010

.015

.030

.035

.040

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.015

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 2.3 M∞ = 2.3

M∞ = 4.0

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 4.0

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

.025

.020

.005

–.005

.005

–.3

–.2

–.1

.1

.2

.3

.4

0

Page 67: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

65

(b) M∞ = 4.63.

Figure 45. Concluded.

–.015

–.010

0

.010

.015

.030

.035

.040

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

1086420–2–4–6–8

M∞ = 4.63 M∞ = 4.63

12α

1086420–2–4–6–8

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°.025

.020

.005

–.005

Page 68: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

66

(a) M∞ = 1.6 and 1.8.

Figure 46. Pitch control effectiveness of elevons for cranked-wing fully integrated configuration with moment referencecenter at 59 percent of body length.

0

.010

.015

.030

.035

.045

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

M∞ = 1.6 M∞ = 1.6

M∞ = 1.8

12α

1086420–2–4–6–8

12α

1086420–2–4–6–8 12α

1086420–2–4–6–8

M∞ = 1.8

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

.025

.020

.005

–.005

.005

–.3

–.2

–.1

.1

.2

.3

.4

0

.040

Page 69: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

67

(b) M∞ = 2.0 and 2.3.

Figure 46. Continued.

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

M∞ = 2.0 M∞ = 2.0

M∞ = 2.3

12α

1086420–2–4–6–8

M∞ = 2.3

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

–.3

–.2

–.1

.1

.2

.3

.4

0

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

12α

1086420–2–4–6–8

.005

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

12α

1086420–2–4–6–8

.005

Page 70: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

68

(c) M∞ = 4.0 and 4.63.

Figure 46. Concluded.

12–.3

–.2

–.1

0

.1

.2

.3

.4

α

Lif

t coe

ffic

ient

, CL

Lif

t coe

ffic

ient

, CL

1086420–2–4–6–8

M∞ = 4.0 M∞ = 4.0

M∞ = 4.63

12α

1086420–2–4–6–8

M∞ = 4.63

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

δA = 0°, δE = 0°

δA = 0°, δE = 20°

δA = 0°, δE = –20°

–.3

–.2

–.1

.1

.2

.3

.4

0

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

12α

1086420–2–4–6–8

.005

–.010

–.005

0

.010

.015

.020

.025

.030

.035

.040

Pitc

hing

-mom

ent c

oeff

icie

nt, C

M

12α

1086420–2–4–6–8

.005

Page 71: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

69

Figure 47. Lateral-directional stability of straight-wing fully integrated configuration.

Figure 48. Lateral-directional stability of cranked-wing fully integrated configuration.

Stable

Stable

–.0016

–.0004

0

.0004

.0008

.0012

.0016

.0020

12–.0002

–.0001

0

.0001

.0002

.0003

.0004

.0005

.0006

.0007

.0008

α

Yaw

ing-

mom

ent d

eriv

ativ

e, C

Rol

ling-

mom

ent d

eriv

ativ

e, C

lβ1086420–2–4–6 12

α1086420–2–4–6

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

–.0008

–.0012

Stable

Stable

–.0028

–.0008

0

.0008

.0012

.0016

.0020

.0024

12–.0002

.00020

.0008

.0010

.0012

.0014

.0016

.0018

.0020

.0022

α

Yaw

ing-

mom

ent d

eriv

ativ

e, C

Rol

ling-

mom

ent d

eriv

ativ

e, C

1086420–2–4–6 12α

1086420–2–4–6

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

–.0016

–.0020

M∞ = 1.6

M∞ = 1.8

M∞ = 2.0

M∞ = 2.3

M∞ = 4.0

M∞ = 4.63

M∞ = 1.6

M∞ = 1.8

M∞ = 2.0

.0006

.0004

.0004

–.0004

–.0012

–.0024

Page 72: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

70

Figure 49. Effects of vertical tail on lateral-directional stability of straight-wing fully integrated configuration atM∞ = 4.0.

Figure 50. Effects of vertical tail on lateral-directional stability of cranked-wing fully integrated configuration atM∞ = 4.0.

Stable Stable

–.0016

–.0004

0

.0004

.0008

.0012

.0016

12–.0006

–.0004

0

.0002

.0004

.0006

.0008

.0010

α

Yaw

ing-

mom

ent d

eriv

ativ

e, C

Rol

ling-

mom

ent d

eriv

ativ

e, C

lβ1086420–2–4–6 12

α1086420–2–4–6

–.0008

–.0012

Vertical tail off

Vertical tail on

Vertical tail off

Vertical tail on

–.0002

StableStable

–.0016

–.0004

0

.0004

.0008

.0012

.0016

12–.0006

–.0004

0

.0002

.0004

.0006

.0008

.0010

α

Yaw

ing-

mom

ent d

eriv

ativ

e, C

Rol

ling-

mom

ent d

eriv

ativ

e, C

1086420–2–4–6 12α

1086420–2–4–6

–.0008

–.0012

Vertical tail off

Vertical tail on

Vertical tail off

Vertical tail on

–.0002

Page 73: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

71

Figure 51. Aileron effectiveness on lateral-directional stability of straight-wing fully integrated configuration;δA = ±20° and δE = 0°.

Figure 52. Aileron effectiveness on lateral-directional stability of cranked-wing fully integrated configuration;δA = ±20° and δE = 0°.

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

–.010

–.004

0

.006

.008

0

.008

.010

.012

.014

.016

Rol

ling

mom

ent,

∆Cl

Yaw

ing

mom

ent,

∆Cn

12α

1086420–2–4–6

–.006

.010

.006

.004

.002

.004

.002

–.002

–.008

–812α

1086420–2–4–6–8

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

–.010

–.004

0

.006

.008

0

.008

.010

.012

.014

.016

Rol

ling

mom

ent,

∆Cl

Yaw

ing

mom

ent,

∆Cn

12α

1086420–2–4–6

–.006

.010

.006

.004

.002

.004

.002

–.002

–.008

–812α

1086420–2–4–6–8

M∞ = 2.0M∞ = 1.8

M∞ = 1.6

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

M∞ = 2.0M∞ = 1.8

M∞ = 1.6

Page 74: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

72

Figure 53. Combined roll/pitch effectiveness on lateral-directional stability of cranked-wing fully integrated configura-tion; δA = ±20° and δE = 20°.

–.010

–.004

0

.006

.008

0

.008

.010

.012

.014

.016

Rol

ling

mom

ent,

∆Cl

Yaw

ing

mom

ent,

∆Cn

12α

1086420–2–4–6

–.006

.006

.004

.002

.004

.002

–.002

–.008

–812α

1086420–2–4–6–8

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

.010

M∞ = 2.0M∞ = 1.8

M∞ = 1.6

M∞ = 4.63M∞ = 4.0

M∞ = 2.3

M∞ = 2.0M∞ = 1.8

M∞ = 1.6

Page 75: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,
Page 76: Aerodynamic Characteristics of Two Waverider-Derived ...mln/ltrs-pdfs/NASA-96-tp3559.pdf · types of missions including hypersonic cruise vehicles, single-stage-to-orbit vehicles,

Form ApprovedOMB No. 0704-0188

Public reporting burden for this collection of information is estimated to average 1 hour per response, including the time for reviewing instructions, searching existing data sources,gathering and maintaining the data needed, and completing and reviewing the collection of information. Send comments regarding this burden estimate or any other aspect of thiscollection of information, including suggestions for reducing this burden, to Washington Headquarters Services, Directorate for Information Operations and Reports, 1215 JeffersonDavis Highway, Suite 1204, Arlington, VA 22202-4302, and to the Office of Management and Budget, Paperwork Reduction Project (0704-0188), Washington, DC 20503.

1. AGENCY USE ONLY (Leave blank) 2. REPORT DATE 3. REPORT TYPE AND DATES COVERED

4. TITLE AND SUBTITLE 5. FUNDING NUMBERS

6. AUTHOR(S)

7. PERFORMING ORGANIZATION NAME(S) AND ADDRESS(ES)

9. SPONSORING/MONITORING AGENCY NAME(S) AND ADDRESS(ES)

11. SUPPLEMENTARY NOTES

8. PERFORMING ORGANIZATIONREPORT NUMBER

10. SPONSORING/MONITORINGAGENCY REPORT NUMBER

12a. DISTRIBUTION/AVAILABILITY STATEMENT 12b. DISTRIBUTION CODE

13. ABSTRACT (Maximum 200 words)

14. SUBJECT TERMS

17. SECURITY CLASSIFICATIONOF REPORT

18. SECURITY CLASSIFICATIONOF THIS PAGE

19. SECURITY CLASSIFICATIONOF ABSTRACT

20. LIMITATIONOF ABSTRACT

15. NUMBER OF PAGES

16. PRICE CODE

NSN 7540-01-280-5500 Standard Form 298 (Rev. 2-89)Prescribed by ANSI Std. Z39-18298-102

REPORT DOCUMENTATION PAGE

July 1996 Technical Paper

Aerodynamic Characteristics of Two Waverider-Derived Hypersonic CruiseConfigurations WU 466-02-01-01

Charles E. Cockrell, Jr., Lawrence D. Huebner, and Dennis B. Finley

L-17479

NASA TP-3559

Cockrell and Huebner: Langley Research Center, Hampton, VA; Finley: Lockheed-Fort Worth Company,Fort Worth, TX.

An evaluation was made on the effects of integrating the required aircraft components with hypersonic high-liftconfigurations known as waveriders to create hypersonic cruise vehicles. Previous studies suggest that waveridersoffer advantages in aerodynamic performance and propulsion/airframe integration (PAI) characteristics over con-ventional non-waverider hypersonic shapes. A wind-tunnel model was developed that integrates vehicle compo-nents, including canopies, engine components, and control surfaces, with two pure waverider shapes, both conical-flow-derived waveriders for a design Mach number of 4.0. Experimental data and limited computational fluiddynamics (CFD) solutions were obtained over a Mach number range of 1.6 to 4.63. The experimental data show thecomponent build-up effects and the aerodynamic characteristics of the fully integrated configurations, includingcontrol surface effectiveness. The aerodynamic performance of the fully integrated configurations is not compara-ble to that of the pure waverider shapes, but is comparable to previously tested hypersonic models. Both configura-tions exhibit good lateral-directional stability characteristics.

Hypersonic cruise; Waveriders; Airbreathing vehicles 73

A04

NASA Langley Research CenterHampton, VA 23681-0001

National Aeronautics and Space AdministrationWashington, DC 20546-0001

Unclassified–UnlimitedSubject Category 02Availability: NASA CASI (301) 621-0390

Unclassified Unclassified Unclassified