cfd a review of out-of-ground-effect propulsion-induced.pdf

Upload: mojicap

Post on 02-Jun-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    1/17

    Progress in Aerospace Sciences 41 (2005) 175191

    A review of out-of-ground-effect propulsion-induced

    interference on STOVL aircraft

    A.J. Saddington, K. Knowles

    Aeromechanical Systems Group, Department of Aerospace, Power and Sensors, Cranfield University, RMCS, Shrivenham, Swindon,

    Wiltshire, SN6 8LA, UK

    Available online 14 June 2005

    Abstract

    This paper provides a review of the out-of-ground-effect propulsion-induced interference on the aerodynamics of jet-

    lift short take-off and vertical landing (STOVL) aircraft. Two main flight regimes are discussed; hover and transition

    wherein the main fluid dynamics phenomena that cause the interference are presented. Where possible, an engineering

    assessment is made of the effect of jet and/or airframe configuration parameters on the observed interference effects.

    r 2005 Elsevier Ltd. All rights reserved.

    Contents

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1762. The free jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

    3. Hover out of ground effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

    3.1. The effect of jet decay rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

    3.2. The effect of nozzle pressure ratio (NPR). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

    3.3. The effect of nozzle length and projection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

    3.4. The effect of planform shape and position . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

    3.5. The effect of off-axis nozzle positions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    4. Transition out of ground effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    4.1. The free jet in a cross-flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    4.1.1. Correlation parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

    4.1.2. Jet trajectory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

    4.2. Jet-induced interference effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1834.2.1. The effect of velocity ratio,Ve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

    4.2.2. The effect of area ratio, AnS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

    4.2.3. The effect of nozzle geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

    4.2.4. The relative location of wing and jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

    www.elsevier.com/locate/paerosci

    ARTICLE IN PRESS

    0376-0421/$ - see front matterr 2005 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.paerosci.2005.03.002

    Corresponding author.

    E-mail addresses: [email protected] (A.J. Saddington), [email protected] (K. Knowles).

    http://www.elsevier.com/locate/paeroscihttp://www.elsevier.com/locate/paerosci
  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    2/17

    4.2.5. The effect of nozzle vector angle, dn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

    4.2.6. Lift improvement devices (LIDs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

    4.3. Jet and intake-induced interference effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

    5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

    1. Introduction

    For over 40 years engineers have struggled with the

    concept of giving a jet-powered fixed-wing aircraft the

    capability to hover and transition to and from wing-

    borne flight. From the tens of designs studied, only

    about 20 jet-powered STOVL aircraft made it to the

    flight test stage and of these, only two (the Harrier and

    Yak-38) became operational, a testimony to the

    difficulty in designing a successful STOVL aircraft.

    With development of the Joint Strike Fighter (JSF)

    nearing completion it is perhaps poignant to reflect on

    the lessons learnt throughout the development of theseaircraft.

    This paper presents a review of the main out-of-

    ground-effect propulsion-induced aerodynamic interfer-

    ence encountered by jet-lift STOVL aircraft. It will draw

    from several earlier surveys [16] as well as presenting

    more recent research results.

    2. The free jet

    Many generalised studies of axisymmetric free jets

    have been made and their characteristics are well

    documented [79]. Studies focussing on particular

    aspects of free jets such as the self-similar region

    downstream of the nozzle [1012], the effect of initial

    conditions[1316], the effect of a co-flowing stream[17]

    and large-scale structures within the jet[18,19]have also

    been made.

    With regard to STOVL aircraft applications high-

    speed (in particular under-expanded) free jets are of

    relevance and these too have been well documented

    [2024]. For a circular convergent nozzle, three varia-

    tions of the flow pattern are possible, depending on the

    nozzle pressure ratio (NPR, the ratio of settling chamber

    total pressure, Pc to ambient static pressure, pa). Thethree variations are the subsonic jet, the under-expanded

    jet and the highly under-expanded jet [23]. Note that

    with a convergent-divergent nozzle an additional flow

    pattern exists, that of the overexpanded jet. This occurs

    when the nozzle is operating off-design at too high a

    NPR[25].

    Jet spreading rate and entrainment characteristics

    have been the subject of a number of studies. Anderson

    and Johns [20] considered a range of jet exit Mach

    numbers from 1.4 to 3.53. The jets were found to have a

    mean cone half angle of 4:6 with this value being

    independent of NPR. Later studies by Moustafa [24]

    Nomenclature

    An nozzle exit area

    dn nozzle diameter

    dne equivalent nozzle diameter

    D angular mean planform diameteri index (see Eq. (3))

    k1, k2 constants in Eqs. (2), (4) and (7)

    L lift

    m constant in Eq. (10)

    _mn nozzle exit mass flow rate

    _mx mass flow rate at a streamwise distance x

    max maximum

    M pitching moment

    Mn nozzle exit Mach number

    M1 freestream (cross-flow) Mach number

    n constant in Eq. (10)

    NPR nozzle pressure ratiopa ambient static pressure

    Pc settling chamber total pressure

    Pertotal total jet perimeter

    qn nozzle exit dynamic pressure

    qx dynamic pressure at a streamwise distancex

    r radius

    rn nozzle radius

    S planform area

    T thrustuc mean jet centreline velocity

    Ve effective velocity ratio (see Eq. (9))

    Vn nozzle exit velocity

    V1 freestream (cross-flow) velocity

    x streamwise distance from nozzle exit

    z distance perpendicular to nozzle exit

    x x=rna;b constants in Eq. (10)dn nozzle vector angle

    D incremental

    y angular measurement (see Eq. (8))

    k constant in Eq. 1re ra=rnrn nozzle exit static density

    ra ambient static density

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191176

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    3/17

    and Lau et al. [26]indicated that jet spreading rate is in

    fact strongly influenced by NPR and jet exit Mach

    number. Nozzle geometry has also been found to

    influence spreading characteristics with elliptical, rec-

    tangular and lobed nozzles all showing greater spreading

    rates than jets from a circular nozzle [27].

    Ricou and Spalding [28] summarise results for the

    mass entrained by a jet, based on measured velocity

    profile data and augmented with analytical solutions,

    showing that for given jet exit conditions air is entrained

    into a jet at a constant rate. Witze [29] making use of

    Kleinsteins [30] theoretical formulation and the results

    of 13 experimental studies, derived an empirical

    formulation (Eq. (1)) for the decay in the centreline

    velocity for varying jet exit Mach numbers.

    uc x 1 exp 1

    k x re0:5 0:7

    !, (1)

    where k is a proportionality constant given by

    k 0:081 0:16Mn re0:22 Mno1,

    k 0:063M2n 10:15 Mn41.

    The influence of temperature on the characteristics of

    free jets has also been widely investigated [29,3133],

    particularly with respect to spreading rates [31] and

    entrainment rates [33].

    3. Hover out of ground effect

    The basic flow-field associated with a jet-lift STOVL

    aircraft hovering out of ground effect is illustrated in

    Fig. 1 [34]. In all cases, the lifting jets issuing from the

    aircraft mix with ambient air to generate a complicated

    three-dimensional flow-field. In general, when one or

    more jet-lift engines are installed in an aircraft fuselage

    or wing, and exhaust vertically downwards, a small lift

    loss is induced by the entrainment action of the jet(s).

    For a simple circular planform with a single central

    round jet having a uniform exit velocity distribution

    (Fig. 2), the reduction in lift is about 2% of the installed

    thrust T, for a nozzle area to planform area ratio

    An=S 0:01 [35,36]. This lift loss reduces steadily, butnon-linearly, as the ratioAn=S increases, becoming onlyabout 0.5% for An=S 0:11 and negligible if An=S isgreater than 0.25[35,36]. The lift losses on the plate arise

    from the entrainment action of the jet, which mixes with

    the surrounding air and sets up a cross-flow over the

    bottom of the plate, thereby inducing a suction pressure.

    The four-jet configuration of Otis[37]indicated that lift

    losses of between 3 and 4 per cent could be encountered

    for typical multiple-jet configurations.

    3.1. The effect of jet decay rate

    Gentry and Margason [38] made lift loss measure-

    ments using a circular and a rectangular plenum

    chamber. The latter was designed to fit inside the

    fuselage of a wind tunnel model and was smaller than

    ideally desired. A comparison of the loads induced on a

    circular plate by a single nozzle fitted to the circular and

    rectangular plenum chambers (Fig. 3(a)), indicated that

    the loads induced on the circular plate mounted on the

    Entrained air

    Intake flow

    Fig. 1. The flow field surrounding a V/STOL aircraft in hover

    [34].

    L

    Fig. 2. The lift loss generated on a flat plate by a single nozzle

    [35].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 177

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    4/17

    rectangular plenum were three to four times as large as

    those induced on the same plate-nozzle arrangement

    with the circular plenum chamber.

    Measurements of the velocity profile obtained from

    the rectangular plenum indicated a distorted velocity

    distribution with a deficit at the centre, whereas the flowfrom the circular plenum had a uniform top hat

    velocity profile. The flow through the rectangular

    plenum was found to be very turbulent and it was

    suspected that the high turbulence levels were causing

    the higher induced loads. To confirm the hypothesis,

    fairings and baffles were added to the rectangular

    plenum chamber to improve the flow quality and a

    strut was inserted in the circular plenum to degrade the

    flow quality. This also produced a much greater

    distortion of the exit velocity profile than was present

    on the original rectangular plenum chamber. Fig. 3(a)

    shows that these changes increased the lift loss for the

    circular plenum chamber and greatly reduced the losses

    for the rectangular plenum chamber.

    The lift losses did not correlate with the exit velocity

    distribution, however, they were found to correlate withthe rate of decay of the jet with distance downstream

    from the nozzle exit (Fig. 3(b)). These curves show the

    peak non-dimensional dynamic pressure in the jet at a

    non-dimensional distance downstream of the nozzle. The

    circular plenum chamber gave the smallest lift loss and

    had the lowest decay rate. The original rectangular

    plenum produced the greatest lift loss and had the highest

    decay rate. The modifications made to the circular and

    rectangular plenum chambers produced jets with similar

    decay rates, giving similar lift losses. The correlation was

    also found to hold true for multiple-jet configurations.

    Fig. 4 compares the lift loss and decay rates for threedifferent jet arrangements (all using the modified

    rectangular plenum chamber). An empirical equation

    was developed to predict the lift loss per unit thrust based

    on the jet decay rate and model geometry[38].

    DL

    T

    ffiffiffiffiffiffiS

    An

    r k1k2, (2)

    where k1 0:009 and

    k2

    ffiffiffiffiffiffiffiffiffiffiffiffiffi ffiffiffiffiffiffiffiffiffiffiffiffi ffiffiffiffiffiffiffiffiffiffiffi

    qqx=qn

    @x=dne

    max

    x=dnei

    vuuut , (3)

    (S/An)

    L/T

    0 1 2 3 4 5 6 7 8 9 10-0.05

    -0.04

    -0.03

    -0.02

    -0.01

    0

    Single jet

    Four jet

    Eight jet

    Four slot jet

    Single jet

    Four jet

    Eight jetFour slot jet

    x/dne

    qx

    /(Pc-pa

    )

    0 1 2 3 4 5 6 7 8 9 100.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    (a)

    (b)

    Fig. 4. (a) Induced load and (b) jet decay for single-jet and

    several multiple-jet configurations NPR 1:89 [38].

    NPR

    L/T

    1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4-0.04

    -0.03

    -0.02

    -0.01

    0

    Cylindrical plenum, clean nozzle

    Cylindrical plenum, restriction in nozzle

    Original rectangular plenum

    Modified rectangular plenum

    Cylindrical plenum, clean nozzle

    Cylindrical plenum, restriction in nozzleOriginal rectangular plenum

    Modified rectangular plenum

    x/dn

    qx

    /(Pc-pa

    )

    0 1 2 3 4 5 6 7 8 9 100.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    (a)

    (b)

    Fig. 3. The effect of plenum chamber and nozzle configuration

    on (a) induced load on several plate shapes and (b) on jet decay

    NPR 1:64 [38].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191178

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    5/17

    wherexis the distance along the nozzle axis, the origin of

    which is at the exit plane. The lift loss is therefore

    proportional to the square root of the maximum rate of

    decay of the jet divided by the distance to the point at

    which this occurs. The correlation showed good agree-

    ment with the test data.

    It is clear that the use of rapid mixing nozzles, if

    installed on a STOVL aircraft, will have a detrimental

    effect on the aircrafts hover performance. These

    nozzles, which are currently being developed for reduced

    radar and infra-red signatures [39] and for minimising

    ground erosion problems, are designed to promote high

    decay rates of the jet potential core and may employ

    techniques including: geometric modifications such as

    lobes, notches or castellations; excrescences such as tabs

    or vortex generators and fluidic techniques such as swirl

    or shear layer excitation. An extensive literature review

    of jet mixing enhancement is given by Wong [40], to

    which the reader is referred.When multiple jets are dispersed over the planform,

    the percentage lift loss is greater than for the corre-

    sponding single jet of the same equivalent diameter

    ratio,dne [35]. This is partly due to the increased mixing

    rate, but also to the additional pressure drop produced

    on the lower surface between the jets because of the

    constricting effect of the jets themselves on the entrained

    flow. This does not appear, however, to be a strong

    effect and the lift loss seems unlikely to exceed 5%

    unless the jets are arranged in rows or elongated narrow

    slots, thus tending to enclose a significant amount of the

    planform area.

    Using a Pratt and Witney J85 engine equipped with

    two alternative jet pipes; one for a single jet investiga-

    tion, and one with a special exhaust ducting to produce a

    four-jet configuration, McLemore [41] ran tests with

    three different sizes of rectangular plates to produce

    three different ratios of planform area to jet area An=S.The comparison of these real engine data with the model

    results of Gentry and Margason [38] shows good

    agreement (Fig. 5).

    Using data from Fleming [42] and Higgins et al.

    [43,44],Kuhn and McKinney[45](and later Hammond

    [46]), present a comparison of experiments conducted

    using jets of different temperature. These suggest thattemperature does not appear to have an appreciable

    effect on the jet decay rate except for a circular nozzle

    (Fig. 6).

    The studies on jet decay rate demonstrate that if

    accurate lift loss data are to be obtained from wind tunnel

    models then the decay rate of the simulated jets should

    accurately match those of the full size installation.

    3.2. The effect of nozzle pressure ratio (NPR)

    Gentry and Margason observed that the measured lift

    losses showed a reduction with increasing NPR for a

    constant nozzle to planform area ratio[38]. At the time a

    theoretical model was not available to explain this. NPR

    effects were looked at, however, in an extensive study of

    jet-induced interference effects undertaken by Shumpert

    and Tibbetts[47]. Tests were conducted on a 14 per cent

    scale model of a jet VTOL aircraft with a fuselage-

    mounted lift system at pressure ratios of 1.4, 1.7, 2.0 and

    2.3. It was discovered that the lift loss did not follow the

    correlation proposed by Gentry and Margason (Eq. (2)).

    The results showed that there was a measurable

    (S/An)

    L/T

    0 1 2 3 4 5 6 7 8 9 10-0.05

    -0.04

    -0.03

    -0.02

    -0.01

    0

    Single jet, small scale

    Four jet, small scale

    Single jet, J85

    Four jet, J85

    Single jet, small scale

    Four jet, small scale

    Single jet, J85

    Four jet, J85

    x/dne

    qx

    /(Pc-pa

    )

    0 1 2 3 4 5 6 7 8 9 10 11 120.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2

    (a)

    (b)

    Fig. 5. Comparison of (a) induced load and (b) jet decay for the

    small-scale results of Ref. [38]and a J85 turbojet [41].

    x/dne

    qx

    /(Pc-pa

    )

    0 2 4 6 8 10 12 14 160.0

    0.2

    0.4

    0.6

    0.8

    1.0

    1.2Circular nozzle (70F)Circular nozzle (1200F)Single slot nozzle (70F)Single slot nozzle (1200F)Four slot nozzle (70F)Four slot nozzle (1200F)

    Fig. 6. The effect of temperature and nozzle configuration on

    jet decay[45].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 179

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    6/17

    discrepancy between the lift losses at different NPRs for

    a constant nozzle area and a modification to Eq. (2) was

    proposed.

    DL

    T

    ffiffiffiffiffiffiS

    An

    r NPR0:64k1k2. (4)

    The constant k1 was found to be aircraft configuration

    dependent and for the tests conducted, varied between

    0:01 and 0:016. The parameter k2 was unchanged(see Eq. (3)). As this correlation suggests, an increase in

    NPR is seen to result in a decrease in percentage lift loss.

    This effect was found to be similar for all configurations.

    The NPR effect described above is explained by the jet

    entrainment rate of different NPR jets. At an axial

    distance, x downstream of the nozzle exit plane, the

    ratio of mass flow rate in the jet to the mass flow rate at

    the nozzle exit is given by[28]

    _mx

    _mn 0:32

    x

    dn

    ffiffiffiffiffira

    rn

    r . (5)

    Eq. (5) shows that as NPR is increased (and hence the jet

    density), the mass flow rate of the jet at a given

    streamwise location is reduced. The decay rate of a high

    NPR jet will, therefore, be less than that of a lower NPR

    jet. Comparison with data collected by Schwantes [48]

    and reported by Ransom and Smy [49] suggests that

    although Eq. (5) gives the correct trend for jet

    entrainment, the magnitude calculated may not be relied

    upon.

    Shumpert and Tibbetts also conducted experiments at

    constant thrust, i.e. the nozzle area was reduced as theNPR was increased. Varying NPR at constant thrust

    had no measurable effect on lift loss [47].

    A more direct, easier to use method for estimating

    hover lift losses is presented by Kotansky[50]. Correla-

    tion of data from various single and multiple jet

    configurations[51]resulted in the following equation:

    DL

    T 0:0002528

    ffiffiffiffiffiffiS

    An

    r

    NPR0:64 Pertotal

    dne

    1:581. 6

    Eq. (6) implicitly accounts for the higher decay rate ofmultiple jet configurations in terms of equivalent nozzle

    diameter but does not account for higher decay rates

    caused by jet exit conditions involving high entrainment

    rates i.e. rapid mixing nozzles. If higher than normal

    turbulence and decay rates are involved, Kotansky

    suggests that Eq. (4) should be used [50].

    3.3. The effect of nozzle length and projection

    Gentry and Margason [38] investigated the effect of

    nozzle length and projection on induced loads for a

    single round jet with the modified rectangular plenum

    chamber described in Section 3.1 and with rectangular-

    planform plates. In these experiments, nozzle length

    was defined as the distance from the bottom of the

    settling chamber (the start of the nozzle profile) to

    the nozzle exit plane. Nozzle projection, by contrast,

    was the distance the nozzle exit plane extended below

    the planform surface. The findings are summarised in

    Table 1for NPRs of 1.5 and 2.0.

    Extending the nozzle length caused a slight reduction

    in the lift loss for the case of zero projection.

    Measurements of the jet decay rate for the different

    nozzle lengths showed that the shortest nozzle had the

    highest decay rate. A further reduction in lift loss was

    obtained by projecting the nozzle through the circular

    plate distances of 0:5dn and 1:0dn. This reduction is dueto the increased distance between the plate and the free

    surface of the jet which is entraining air.

    3.4. The effect of planform shape and position

    In order to determine the effect of planform shape on

    the induced lift loss, Gentry and Margason tested four

    plate shapes on their circular plenum chamber [38]. The

    four shapes were circular, square, rectangular (aspect

    ratio 1.52) and triangular (equilateral). In all cases the

    ratio of nozzle exit area to plate area, An=Swas 0.0144and the centre of the nozzle was located at the centroid

    of the plate. There was no measurable effect of planform

    shape on the lift loss. The effect of wing height was also

    investigated. As would be expected, a low wing

    configuration had more lift loss than a mid-mountedwing which in turn performed worse than a high wing.

    This is due to the relative proximity of the wing to the

    mixing region of the jet.

    The effect of planform lip shape was investigated by

    Ing and Zhang [52]. An experiment was carried out to

    study the parameters affecting single-jet ground-effect

    hover lift loss and to identify the cause behind the large

    discrepancies in lift loss between the experiments of

    Wyatt[53]and Corsiglia et al.[54],commonly known as

    Table 1Lift loss for various nozzle lengths and projections, An=S 0:0105[38]

    Nozzle length

    dn

    Nozzle

    projection

    dn

    DL=TNPR 1:5

    DL=TNPR 2:0

    0.248 0.0 0.0187 0.0175

    0.910 0.0 0.0163 0.0155

    1.410 0.0 0.0163 0.0148

    0.910 0.5 0.0130 0.0129

    1.410 0.5 0.0133 0.0126

    1.410 1.0 0.0116 0.0107

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191180

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    7/17

    the Wyatt anomaly. The cause of the discrepancies was

    traced to a single geometrical parameter, namely the

    planform plate edge geometry, which significantly

    affected the flow separation underneath the plate. The

    out of ground effect lift loss was, however, found to be

    insensitive to the edge geometry.

    3.5. The effect of off-axis nozzle positions

    The effect of positioning a nozzle away from the

    fuselage centre-line was part of a study carried out by

    Welte[55]. Tests on a model of the VTOL transporter,

    Dornier Do-31, indicated that a centrally located jet will

    produce more lift loss than a peripherally located jet.

    The podded lift engines installed at the wing tips induced

    a lift loss of 2.2 per cent compared with 3.6 per cent for

    the lift/cruise engine installed below the inner wing

    section. Welte postulated that the constant k1 in Eq. (3)

    is universal and an additional term could be introducedinto the equation to take into account configuration

    effects. The following equation was developed [55].

    DL

    T

    D

    dneNPR0:64k1k2, (7)

    where D is the angular mean planform diameter and is

    given by

    D 1

    p

    Z 2p0

    ry dne

    2

    dy, (8)

    where k1 0:009 and k2 is unchanged from Eq. (3).

    It was found that for the lift/cruise engines, both Eqs.(4) and (7) predicted the lift loss quite well, 4.0 and 3.4

    per cent, respectively. With the peripherally located

    lift engines at the wing tip, Eq. (4) failed to produce

    a satisfactory result, again predicting 4.0 per cent

    (as would be expected) compared with 2.4 per cent for

    Eq. (7).

    4. Transition out of ground effect

    The transition from hover to wing-borne flight is a

    very important design consideration for STOVL air-craft. Fig. 7 shows the principle aerodynamic phenom-

    ena affecting a STOVL aircraft in transition out of

    ground effect [56]. During the transition from hovering

    to conventional flight, the efflux from the lifting jet(s) is

    swept back by the freestream and rolled up into pairs of

    vortices. These vortices, along with the entrainment

    action and blockage effect of the jet(s), induce suction

    pressures on the bottom of the configuration beside and

    behind the jet(s) and a smaller region of positive

    pressure ahead of the jet(s). In most cases, these induced

    pressures result in a nose-up pitching moment and a loss

    of lift on the aircraft.

    4.1. The free jet in a cross-flow

    The jet in cross-flow has received considerable

    research attention because it is a fundamental fluid

    dynamic phenomena with many applications such as

    smoke plumes from chimneys, the dispersal of liquids in

    streams, jet engine combustors and reaction control jets

    on rockets and missiles. The most widely studied, and

    reviewed, application, however, is related to STOVL

    aircraft [57,58].

    A jet exhausting into a cross-flow generates a complex

    flow-field with several distinguishable features. When

    the jet efflux exits the nozzle it is deflected by the

    freestream to follow a curved path downstream while its

    cross-section changes. For the case of a circular jet, there

    are stagnation points upstream and downstream and

    minimum pressures at the lateral edges. As a conse-

    quence, the jet spreads laterally into an oval shape. At

    the same time the cross-flow shears the jet fluid along the

    lateral edges downstream to form a kidney-shaped

    cross-section. At increasing distances along the jet path

    this shearing folds the downstream face over itself to

    form a vortex pair which dominates the flow. Associated

    with the vortex pair is the flow induced into the wake

    region of the jet from the freestream. Fig. 8 shows the

    topology of a jet in cross-flow [58].

    4.1.1. Correlation parameters

    Early jet in cross-flow investigations usually charac-

    terised the test condition by the ratio of freestream

    velocity to jet exit velocity, V1=Vj. It was soonrecognised, however, that this ratio was not appropriate

    for hot jets. Williams and Wood observed that non-

    dimensional increments in lift, drag and pitching

    moment were primarily, but not solely, a function of

    jet to freestream velocity ratio [35]. The influence of

    jet Reynolds number (based on nozzle diameter) and

    model Reynolds number (based on wing chord) on jet

    Viscous effects onwing and controlsurfaces

    Intake mass flow

    Jet blockage and entrainment

    Fig. 7. The flow field about a STOVL aircraft in transition out

    of ground effect[56].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 181

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    8/17

    interference effects were secondary. They showed that

    the influence of temperature and compressibility effects

    could better be accounted for by the use of an effective

    velocity ratio (Eq. (9)), which is still widely used.

    Ve

    V1

    Vnffiffiffiffiffirarn

    r

    M1

    Mnffiffiffiffiffipa

    pnr

    . (9)

    For STOVL applications the magnitude of Ve for

    transition to wing-borne flight depends on the pressure

    ratio of the lifting propulsion device. For high-pressure

    ratio jet engines, transition will occur with Veo0:2; forthe Harrier, transition will occur with Veo0:3; and forlift fans with low pressure ratios, transition will occur

    with Veo0:5[58].

    4.1.2. Jet trajectory

    One property of jets in cross-flow which has received a

    lot of research attention is the jet trajectory[5963]. The

    most common parameter for defining the trajectory is

    the locus of maximum velocity; however, other defini-

    tions are sometimes used depending on the preference of

    the researcher. These trajectory definitions include: the

    locus of maximum stagnation pressure; the locus of

    maximum dynamic pressure; the locus of maximum

    stagnation temperature and the line of maximum

    vorticity. Since none of these definitions will give the

    same trajectory, it is necessary to exercise great care

    when comparing data from different sources.

    Nearly all the work done to establish a jet trajectory

    has used a subsonic jet in a subsonic cross-flow. A

    widely used empirical equation for the jet centreline

    trajectory is [58],

    x

    dn aVne

    z

    dn m

    z

    dn b cotdn. (10)

    Thez and x origins are taken as the centre of the nozzle

    exit. When the nozzle vector angle is 90 (i.e. perpendi-

    cular to the cross-flow), the second term in Eq. (10)

    becomes zero. The empirical approaches correlate the

    experimental data using similarity laws to obtain the jet

    trajectory. Table 2 summarises the coefficients and

    exponents for jets with a 90 vector angle [6467] and

    for jets with varying vector angles[6870].

    Eq. (10) has been plotted using an effective velocity

    ratio,Ve of 0.25 which covers the range of applicability

    for most of the researchers. Fig. 9 shows the jettrajectories for a nozzle vector angle of 90 while

    Fig. 10shows the trajectories for a vector angle of 60.

    The equations derived by the researchers show good

    agreement with each other. There are slight variations in

    the initial depth of penetration and the rate of jet

    deflection. These can probably be accounted for by

    variations in the initial jet and cross-flow conditions in

    the experiments from which the empirical equations are

    derived. In particular the turbulence intensities of the jet

    and cross-flow, which will affect the mixing rate between

    the two, will have a strong influence on the jet

    Table 2

    Jet trajectory equation terms and applicability

    Author(s) a n m b Ve dj (deg)

    Bradbury [64] 1.08 2.70 3.00 n/a 0.080.43 90

    Kamotani and Greber[65] 1.38 2.61 2.78 n/a 0.130.26 90

    Chaissaing et al. [66] 0:59 Ve2:6 2.60 2.60 n/a 0.160.42 90

    Fearn and Weston[67] 1.08 2.68 2.95 n/a 0.100.33 90

    Ivanov[68] 1.00 2.60 3.00 1.00 0.030.71 60120

    Shandorov [69] 1.00 2.00 2.55 1 Ve2 0.210.71 4590

    Margason[70] 1=4sin2dj 2.00 3.00 1.00 0.100.85 30180

    V

    Jet CL

    Vortex curve

    Separation line

    Vortex pair

    Horseshoe vortex

    Wake vortex street

    Fig. 8. Sketch of the vortex systems associated with a jet in

    cross-flow[58].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191182

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    9/17

    penetration and curvature. Zhang and Hurst [71] have

    observed that with supersonic jets, where the turbulence

    intensity in the jet may be high (due to the breakdown of

    the shock structure), the jet centreline trajectory will

    ultimately lead to less penetration than for a subsonic

    jet.

    Integral methods, e.g. [7277] for jet in cross-flow

    research have also been developed which consider the

    deflecting mechanisms of a jet by taking into account

    either a pressure force or cross-flow momentum

    entrainment or both. More recently the use of computa-

    tional fluid dynamics (CFD) has given researchers anaddition numerical tool, e.g. [7882].

    4.2. Jet-induced interference effects

    In general, the jet-induced interference effects in

    transition are characterised by a loss in lift and a nose-

    up pitching moment.Fig. 11(a) illustrates typical trends

    for a tail-off configuration[83]and shows the lift divided

    by the thrust as a function of angle of attack. Fig. 11(b)

    shows pitching moment as a function of angle of attack.

    In hover, the jets produce a lift which is equal to the net

    thrust. With forward velocity the wing also develops lift

    and, in the absence of interference effects, lift from these

    two sources could be added together to produce the

    solid curve. The jet-induced effects, however, cause a

    loss of lift and the actual lift measured in a wind tunnel

    is shown by the symbols. The difference between the

    calculated and measured curves is the interference lift

    loss, which is generally independent of angle of attack.

    The rolling up of the jet wakes into contra-rotating

    vortices is the primary cause of the interference effects in

    transitional flight[83]. The vortices change the flow field

    in the near-field and induce additional suction pressures

    on the lower surface of the fuselage and wing of the

    aircraft.

    The interaction between the jet and the cross-flow

    means that positive pressures are generated on surfaces

    ahead of the jet and negative pressures behind [84].

    Spreeman[85]measured negative pressures as high as 3

    to 4 times the freestream dynamic pressure. Thepressures diminished with distance from the jet but

    extended 10 to 15 jet diameters downstream and 5 to 10

    (deg.)

    L/T

    -10 0 10 20 300.0

    0.5

    1.0

    1.5

    Calculated

    Measured

    Jet lift

    Power-off lift

    Interference lift loss (L)

    (deg.)

    M

    -10 0 10 20 30

    0

    Interference moment (M)

    Measured power-on moment

    Power-off moment

    (a)

    (b)

    Fig. 11. Jet interference in transition flight (tail off) [83]. (a)

    Lift, (b) pitching moment.

    x/dn

    z/dn

    0 5 10 15 200

    5

    10

    Ivanov [68]

    Shandorov [69]

    Margason [70]

    Fig. 10. Comparison of predicted jet in cross-flow trajectories

    dj 60; Ve 0:25.

    x/dn

    z/dn

    0 5 10 15 200

    5

    10

    Kamotani and Greber [65]Chaissang et.al [66]

    Fearn and Weston [67]

    Bradbury [64]

    Fig. 9. Comparison of predicted jet in cross-flow trajectories

    dn 90; Ve 0:25.

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 183

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    10/17

    diameters to each side of the jet[85].The combination of

    positive pressures ahead of the jet and negative pressures

    behind gives a nose-up pitching moment.

    The following sections aim to describe parameters

    which have been observed to influence the jet-induced

    interference effects. Due to the complex nature of the

    flow fields surrounding STOVL aircraft in transition, it

    is difficult to isolate the effects which different para-

    meters may have on any particular configuration and as

    such only general trends can be identified.

    4.2.1. The effect of velocity ratio, VeThe lift loss on a typical STOVL aircraft in hover has

    already been discussed and is of the order of 2 to 3 per

    cent of the installed thrust. Early experiments conducted

    by Williams and Wood [35] on a simple rectangular-

    wing model with a centrally located jet established that

    the lift increment, DL=T falls steadily below its static

    value as the effective velocity ratio Ve is raised fromzero. The lift loss was accompanied by the expected

    nose-up pitching moments, again increasing as Ve was

    raised. The observed interference effects are due partly

    to the steady growth of the downward load from jet

    interference on the lower surface behind the jet and

    partly from the rearward movement of the centre of this

    jet interference load. Additionally, the normal compo-

    nent of thrust will be reduced because of the jet

    deflection.

    As the velocity ratio is increased further, the lift loss

    reduces but the nose-up pitching moment continues to

    increase. This may be due to two main factors; firstly,

    the configuration being tested may be generating

    aerodynamic lift which will offset the lift loss, and

    secondly, the very high jet deflection angles involved

    may put the jet interference close to the trailing edge of

    the wing. The jets may then act as a crude jet-flap on the

    wing lower surface and thereby generate extra lift by

    super-circulation.

    For practical STOVL aircraft, effective velocity ratios

    below about 0.3 are of primary concern during

    transition with jet-lift schemes although values as high

    as 0.5 could be of importance with high by-pass ratio

    engines or lift fan arrangements. Fig. 12 shows typical

    results for a four-jet configuration[37].

    4.2.2. The effect of area ratio, An=SThe interference lift loss is strongly sensitive to the

    ratio of nozzle area to planform area, An=S. With theconfiguration mentioned above, Williams and Wood

    observed that the lift loss increases as the area ratio is

    decreased and the relationship is approximately linear

    for a simple wing/jet geometry [35]. A larger planform

    provides more area over which the interference pressures

    can act thereby increasing the lift loss. At very low area

    ratios, An=So0:0016, L=T actually fell to zero. For

    more practical area ratios (between 0.01 and 0.04 for a

    typical STOVL aircraft) the lift loss and accompanying

    nose-up moments are less severe, a maximum lift loss of

    25 per cent was recorded with an area ratio of 0.01.

    4.2.3. The effect of nozzle geometry

    Extensive studies of the effect of nozzle geometry were

    carried out by Volger[86].A VTOL model was equipped

    with various arrangements of interchangeable single or

    multiple, round or slotted jets, with and without jet

    deflection. All configurations showed interference lift

    losses that increased with velocity ratio. For most jetarrangements, nose-up pitching moments due to jet

    interference occurred and increased with velocity ratio.

    The configurations which suffered the least interference

    effect on the lift and pitching moment were the single

    longitudinal slot jets, which was thought to be due to

    their more streamlined shape in cross-section and the

    smaller planform area behind the jet.

    An ejector powered STOVL model, equipped with

    various arrangements of slot nozzles, was tested by

    Volger and Goodson [87]. The aerodynamic lift of the

    fuselage, which was almost square in cross-section, at 0

    angle of attack was zero. As velocity ratio was increased

    Ve

    M/Tdne

    0 0.1 0.2 0.30

    0.2

    0.4

    Ve

    L/T

    0 0.1 0.2 0.3-0.4

    -0.2

    0

    (a)

    (b)

    Fig. 12. The effect of wing planform on lift and pitching

    moment S=An 38 [37]. (a) Pitching moment, (b) lift.

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191184

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    11/17

    from 0 (hover) to 0.4, all configurations showed a

    reduction in the combined aerodynamic lift plus thrust.

    Moving the slots outward reduced the affected model

    area directly behind the jets, thereby reducing the

    interference effects between the model, jets and free-

    stream. Yawing the slots either inward or outward at the

    forward end resulted in a reduction in lift and an

    increase in nose-up pitching moments compared with

    the unyawed slots. Yawing the slots outward was less

    detrimental than inward. The effect of splaying the jet

    20 away from the plane of symmetry produced some lift

    improvement and a more nose-down pitching moment

    as a result of less interference between the freestream

    and the outward spreading jets. Any indicated benefit

    from splaying the jet would, however, be partially offset

    by the 6 per cent lift loss resulting from the inclination of

    the thrust axis. A combination of 10 outward yaw of

    the slots and 10 outward splay gave less lift loss than

    configurations with or without angular deflections. Yawappeared to affect the results more than inclination.

    4.2.4. The relative location of wing and jet

    As regards jet position relative to the wing, rearward

    location of the jet exit naturally tends to alleviate the lift

    loss and the nose-up pitching moments arising from jet

    interference, since the surface extent aft of the jet exit is

    reduced. Williams and Wood found evidence to suggest

    that with multiple jets the velocity ratio range over

    which the lift loss occurs can be much reduced and the

    subsequent lift augmentation much increased by the

    adoption of a span-wise row arrangement towards therear of the wing instead of a close cluster [35].

    A generalised study of jet positions from several wing

    chord lengths ahead, to several wing chord lengths

    behind, an unswept wing was reported by Hammond

    and McLemore [88]. In this investigation, an aspect

    ratio 6, unswept, untapered wing-fuselage model

    equipped with a 30 per cent chord slotted Fowler flap

    was used. Two jets, one on either side of the fuselage,

    were positioned at about 25 per cent semi-span and at

    the various longitudinal and vertical positions shown by

    the plus marks in Fig. 13. The jets were mounted

    independently of the wing so that only the aerodynamic

    forces and interference effects were measured on the

    wing. The data show that with the jet exits on the wing

    chord plane, considerable jet interference was experi-

    enced even with the jet as far as four wing chords ahead

    of the wing. Favourable interference effects, however,

    were encountered with the jets beneath and behind the

    50 per cent chord point of the wing and the interference

    effects are most favourable for positions closest to the

    flap. These favourable interference increments were

    believed to be due to the action of the jet in helping

    the flap achieve its full lift potential. Another slightly

    different configuration [88], this time with jets operat-ing simultaneously both in front of and behind the

    wing, indicated an overall favourable interference lift

    effect, again indicating the importance of configuration

    geometry on the jet interference lift and moment

    characteristics. These findings agree well with those of

    Margason and Gentry[112], Schultz and Viehweger[89]

    and Nangia[90].

    Winston et al. [91] report on a configuration which

    was tested to determine the optimum jet exit location

    from an induced lift standpoint. The model was tested

    with the lift/cruise jets in three longitudinal positions as

    shown in Fig. 14 and illustrate the lift improvement

    obtained by successive aft movement of the jet exits

    towards the wing trailing edge.

    The Harrier II, with its new wing and large slotted

    trailing edge flap, initially suffered from interference

    problems whereby the forward nozzle flow interacted

    x/chord

    L/T

    -4 -3 -2 -1 0 1 2 3-0.6

    -0.4

    -0.2

    0

    0.2

    + + +++ + ++ + + + +

    + + + + + +

    + + + ++++ + +

    n = 90 Ve = 0.18

    Fig. 13. The effect of varying chordwise jet location relative to a nearby wing on the induced lift [88].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 185

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    12/17

    with the wing undersurface and inboard pylon at

    transitional forward speeds causing a flow separation

    upstream of the flap. Removal of the scarf directed the

    forward jet to a position further below the wing and

    produced a lift coefficient increment of approximately

    0.2[50].

    4.2.5. The effect of nozzle vector angle,dnWind tunnel tests carried out on a one-sixth scale

    model of the Kestrel (XV-6A) measured the effect of

    vectoring the nozzles on the incremental lift and pitching

    moment at a range of velocity ratios and angles of attack

    [92]. At 0 angle of attack, increasing nozzle vector angle

    to a more vertical position, increased the lift loss and

    nose-up pitching moment. This correlates well with

    similar works[93,94].

    4.2.6. Lift improvement devices (LIDs)

    Spreeman [85] noted that in one full-scale flight

    investigation, deflecting the trailing-edge flaps reduced

    the losses in lift and nose-up pitching moments. The

    beneficial effect of the flaps on this configuration can be

    attributed to positive pressures being built up in front ofthe flap on the lower surface.

    The alleviation of unfavourable jet interference effects

    was shown possible by Williams and Wood with the

    addition of stream-wise fences to the lower surface of

    the airframe, along the sides of the jet exit [35]. The

    geometry of the fence was found to be important. The

    depth of the fence should be at least one jet diameter and

    the stream-wise length at least two jet diameters so as to

    extend forward of the jet exit as well as bounding the

    sides. Flow visualisation suggested that the fences

    reduced the initial curvature of the jet, thus increasing

    the penetration of the jet plume and delaying the growth

    of the trailing vortex flows. This was consistent with the

    apparent negligible effect of the fences at low velocity

    ratios, where the rate of deflection of the jet is inherently

    small. The effect of the fences on a simple fuselage

    model was excellent, reducing maximum transitional lift

    loss from 27% to 7%.

    4.3. Jet and intake-induced interference effects

    During wind tunnel testing of jet-lift STOVL aircraft,

    it is usual to simulate the jet efflux but not the intake

    flows. The intakes, which are commonly faired over or

    are unpowered, are generally tested in separate wind

    tunnel experiments. This may lead to a lack of

    integration between intake [95,96] and nozzle design

    [97]. Separate jet and intake testing ignores any mutual

    interferences between them and airframe, but is adopted

    due to the complexity of building a fully powered, metric

    STOVL model[49,98]. This is in addition to the complextunnel interference problems encountered during transi-

    tion testing [99] necessitating careful planning of test

    procedures [100,101].

    Early research using simple wind tunnel models

    showed only minor interference effects due to intake

    flows [45,112,102,103],which were easily accounted for.

    This lead researchers to conclude that when testing

    models in the transition phase of flight the intake

    interference effects were generally not sufficiently

    important to warrant modelling the intake flows [64].

    Nevertheless some research projects have shown sig-

    nificant jet/intake interference effects. An investigation

    carried out by Mineck and Margason[103], and Mineck

    and Schwendemann [104]determined some interference

    effects between the jets and intakes on a STOVL

    aircraft. Comparison of the force and moment data

    for the intake open and intake closed (using elliptical

    plugs) configurations, indicated that closing the intakes

    decreased the aerodynamic angle of attack by approxi-

    mately 2. Data for the VAK 191B described by

    Haftmann [105] showed a consistent discrepancy be-

    tween the lift loss measured in flight and the lift loss

    measured on a wind tunnel model. The full-size aircraft

    had a transitional lift loss 410% greater than the wind

    tunnel model and a nose-up pitching moment 58%higher. This was probably due, in part, to the use of a

    free flow intake on the model with no suction simulated.

    As more complex STOVL aircraft designs were

    developed, it became increasingly apparent that the

    argument for separate jet and intake testing on STOVL

    aircraft was not valid. The term close-coupled aero-

    dynamics has become increasingly popular in describing

    the engine/airframe designs and operating conditions in

    which the air intake and nozzle system flows create

    significant interferences with respect to each other and/

    or which, when combined, create interferences to

    airframe surfaces differing from the sum of the

    Ve

    L/T

    0 0.1 0.2 0.3 0.4 0.50

    1

    2

    3Calculated (no interference)

    Fig. 14. The effect of chordwise location of deflected lift/cruise

    jets relative to the wing trailing edge on the total lift for a

    supersonic combat lift plus lift/cruise V/STOL aircraft config-

    uration[91].

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191186

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    13/17

    individual interferences. Harris et al. [106] suggest that

    aircraft designs, in which the distance from the intake

    face to the front nozzle relative to the diameter of the

    compressor face is less than 3, will undoubtedly feature

    close-coupled aerodynamics. Whereas for designs in

    which this ratio is greater than 8 the aerodynamics can

    generally be considered to be uncoupled.

    Recent research has concluded that during low speed

    transition Veo0:1 mutual jet/intake interference ef-fects do exist for close-coupled propulsion systems

    [107109]. A generic jet-lift STOVL aircraft equipped

    with powered intakes was tested in the RMCS low-speed

    wind tunnel (Fig. 15). The model was not fully metric

    and so airloads were inferred from static pressure

    tappings on the wing. The model was tested with jet

    and intakes powered separately and simultaneously

    the difference between the airloads giving an indication

    of the mutual interference effect. This is shown in

    Fig. 16, which illustrates the mutual interference as anincrement of wing root sectional lift coefficient for the

    rearmost nozzle position.

    One would expect that the solution to the complexities

    of building a STOVL wind tunnel model would be to use

    computational fluid dynamics (CFD) tools. The combi-

    nation of viscous interactions, broad range of flow

    velocities and extensive regions of separated flow has

    meant that CFD methods tried so far have been unable

    to adequately predict overall airframe forces and

    moments [110]. Although CFD is undoubtedly useful

    in some aspects of STOVL aircraft development,

    designers are still reliant on exhaustive wind tunnel

    tests backed up by large historical databases. Even in

    the 21st Century semi-empirical correlations are still

    being developed for propulsion-induced interference

    effects [111].

    5. Conclusions

    This paper has provided a review of the out-of-

    ground-effect propulsion-induced interference on the

    aerodynamics of jet-lift short take-off and vertical

    landing (STOVL) aircraft. Two main operational

    environments encountered by STOVL aircraft have

    been discussed: hovering out of ground effect and

    transition out of ground effect. The former has included

    the fluid mechanics of an isolated free jet and the

    interference effects caused by such a jet when combined

    with simple and more complex fuselage/wing arrange-

    ments. On the subject of transition out of ground effect,

    the discussion has followed a similar format. Where

    possible, an engineering assessment is made of the effect

    of jet and/or airframe configuration parameters on the

    observed interference effects.

    Although some parametric trends are observed in

    simplified experiments, jet-induced interference effects

    for a complete airframe are difficult to predict and must

    therefore be assessed through rigourous tests of the

    chosen configuration. The evaluation process is still

    predominantly experimental since computational fluid

    dynamics models of complete STOVL aircraft do not yetexist with sufficient fidelity to provide a reliable

    prediction of overall airframe forces and moments.

    References

    [1] Margason RJ. Review of propulsion-induced effects

    on aerodynamics of jet/STOL aircraft, Technical Note

    D-5617, NASA, 1970.

    [2] Barche J. Jet-lift problems of V/STOL aircraft. In:

    AGARD conference proceedings CP-143, fluid dynamics

    panel symposium, Delft, Netherlands, 1974. p. 16.118.

    Ve

    Cl(root)

    0 0.02 0.04 0.06 0.08 0.1 0.12-0.08

    -0.06

    -0.04

    -0.02

    0

    NPR=1.586NPR=2NPR=3NPR=4

    Fig. 16. Mutual interference between propulsive jet and intake

    flows shown as an incremental change in wing root sectional lift

    coefficientdn 60.

    Fig. 15. Front view of the jet/intake model in the RMCS

    1:5 m 1:1 m open-jet wind tunnel: intake suction pipe is seenbehind and to the right of the model; compressed air supply is

    via the light-coloured pipe above the fuselage.

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 187

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    14/17

    [3] Knott PG, Hargreaves JJ. A review of the lifting

    characteristics of some jet-lift V/STOL configurations.

    In: AGARD conference proceedings CP-143, fluid

    dynamics panel symposium, Delft, Netherlands, 1974.

    p. 24.112.

    [4] Leyland DC. Aerodynamics of V/STOL aircraftper-

    formance assessment. AGARD Report R-710, 1984.p. 11.126.

    [5] Kuhn RE. The induced aerodynamics of jet and fan

    powered V/STOL aircraft. In: Krothapalli A, Smith CA,

    editors. Recent advances in aerodynamics, New York :

    Springer; 1986. p. 33773.

    [6] Margason RJ. Propulsion-induced effects caused by out-

    of-ground effects. In: Proceedings of the international

    powered lift conference, Santa Clara, CA, USA, 1987.

    p. 3157, SAE 872307.

    [7] Abramovich GN. The theory of turbulent jets. Cambridge:

    MIT Press; 1963.

    [8] Schlichting H. Boundary-layer theory, 6th ed. New York:

    McGraw-Hill Book Co.; 1968.

    [9] Rajaratnam N. Turbulent jets. Amsterdam: Elsevier;

    1976.

    [10] Gelb GH, Martin WA. An experimental investigation of

    the flow field about a subsonic jet exhausting into a

    quiescent and a low velocity air stream. Can Aeronaut

    Space J 1966;12:33342.

    [11] Wygnanski I, Fiedler H. Some measurements in the self-

    preserving jet. J Fluid Mech 1969;38(3):577612.

    [12] Agrawal A, Prasad A. Properties of vortices in the self-

    similar turbulent jet. Exp Fluids 2002;33(3):56577.

    [13] Oosthuizen PH. An experimental study of low Reynolds

    number turbulent circular jet flow. In: Proceedings of the

    ASME applied mechanics, bioengineering and fluids

    engineering conference, Houston, TX, USA, 1983, ASME83-FE-36.

    [14] Otu gen MV, Kim S, Vradis GC. The near field structure

    of initially asymmetric jets. In: Proceedings of the 34th

    AIAA aerospace sciences meeting and exhibit, Reno, NV,

    USA, 1996, AIAA 96-0202.

    [15] Antonia RA, Zhao Q. Effect of initial conditions on a

    circular jet. Exp Fluids 2001;31(3):31923.

    [16] Xu G, Antonia RA. Effect of different initial conditions

    on a turbulent round free jet. Exp Fluids 2002;33(5):

    67783.

    [17] Antonia RA, Bilger RW. An experimental investigation

    of an axisymmetric jet in a co-flowing air stream. J Fluid

    Mech 1973;61:80522.

    [18] Champagne FH, Crow SC. Orderly structure in jet

    turbulence. J Fluid Mech 1971;48:54791.

    [19] Dimotakis PE, Maike-Lye RC, Papantoniou DA. Struc-

    ture and dynamics of round turbulent jets. Physics of

    Fluids 1983;26:318592.

    [20] Anderson AR, Johns FR. Characteristics of free super-

    sonic jet exhausting into quiescent air. J Jet Propulsion

    1955;25(1):135.

    [21] Love ES, Grigsby CE, Lee LP, Woodling MJ. Experi-

    mental and theoretical studies of axisymmetric free jets.

    Technical Report R-6, NASA, 1958.

    [22] Adamson Jr TC, Nicholls JA. On the structure of

    jets from highly underexpanded nozzles into still air.

    J Aerospace Sci 1959;26(16):1624.

    [23] Donaldson CD, Snedeker RS. A study of free jet

    impingement. Part 1. Mean properties of free and

    impinging jets. J Fluid Mech 1971;45(2):281319.

    [24] Moustafa GH. Further study of high speed single free

    jets. Aeronaut J 1993;97(965):1716.

    [25] Anderson Jr JD. Modern compressible flow. McGraw

    Hill: London; 1990.[26] Lau JC, Morris PJ, Fisher MJ. Measurements in subsonic

    and supersonic free jets using a laser velocimeter. J Fluid

    Mech 1979;93:127.

    [27] Zaman KBMQ. Spreading characteristics of compressible

    jets from nozzles of various geometries. J Fluid Mech

    1999;383:197228.

    [28] Ricou FP, Spalding DB. Measurements of entrainment

    by axisymmetrical turbulent jets. J Fluid Mech 1961;

    11(1):2132.

    [29] Witze PO. Centreline velocity decay of compressible free

    jets. AIAA J 1974;12(4):4178.

    [30] Kleinstein G. Mixing in turbulent axially symmetric free

    jets. J Spacecraft Rockets 1964;1:4038.[31] Lau JC. Effects of exit Mach number and temperature

    on mean flow and turbulence characteristics in round jets.

    J Fluid Mech 1981;105:193218.

    [32] Lepicovsky J. Total temperature effects on centreline

    Mach number characteristics of free jets. AIAA J 1989;

    28(1):47882.

    [33] Curtis P. Investigation into the behaviour of a single

    jet in free air and impinging perpendicularly on

    the ground. Technical Report BAe-KAD-R-RES-3349,

    British Aerospace, 1987.

    [34] Milford CM. Hot gas ingestion in V/STOL. In: Proceed-

    ings of the international powered lift conference, Santa

    Clara, CA, USA, 1987. p. 1729.

    [35] Williams J, Wood MN, Aerodynamic interference effects

    with jet-lift V/STOL aircraft under static and forward

    speed conditions. Technical Report 66403, RAE 1966.

    [36] Wood MN, Jet V/STOL aircraft aerodynamics. In:

    Proceedings of the international congress on subsonic

    aerodynamics, New York, NY, USA, 1967. p. 893920.

    [37] Otis Jr JH. Induced interference effects on a four-jet

    VTOL configuration with various wing planforms in the

    transition speed range. Technical Note D-1400, NASA,

    1962.

    [38] Gentry Jr GL, Margason RJ. Jet-induced lift losses on

    VTOL configurations hovering in and out of ground

    effect. Technical Note D-3166, NASA, 1966.

    [39] Rao GA, Mahulikar SP. Integrated review of stealthtechnology and its role in airpower. Aeronaut J 2002;

    106(1066):62941.

    [40] Wong RYT. Enhancement of supersonic jet mixing.

    Ph.D. thesis, Cranfield University, RMCS, Shrivenham,

    UK, 2000.

    [41] McLemore HC. Jet-induced lift loss of jet VTOL

    configurations in hovering condition. Technical Note

    D-3435, NASA, 1966.

    [42] Fleming WA. Characteristics of a hot jet discharge from a

    jet propulsion engine. Research Memorandum E6L27a,

    NACA, 1946.

    [43] Higgins CC, Wainwright TW. Dynamic pressure and

    thrust characteristics of cold jets discharging from several

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191188

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    15/17

    exhaust nozzles designed for VTOL downwash suppres-

    sion. Technical Note D-2263, NASA, 1964.

    [44] Higgins CC, Kelly DP, Wainwright TW. Exhaust jet

    wake and thrust characteristics of several nozzles

    designed for VTOL downwash suppressiontests in

    and out of ground effect with 70 F and 1200 F nozzle

    discharge temperatures, Contractor Report CR-373,NASA, 1966.

    [45] Kuhn RE, McKinney Jr MO. NASA research on the

    aerodynamics of jet VTOL engine installations. In:

    AGARDograph 103, Specialists Meeting, Tennessee,

    TN, USA, 1965. p. 689713.

    [46] Hammond AD. Thrust losses in hovering for jet VTOL

    aircraft. In: NASA special proceedings, SP-116, 1966.

    p. 16375.

    [47] Shumpert PK, Tibbetts JG. Model tests of jet-induced lift

    effects on a VTOL aircraft in hover. Contractor Report

    1297, NASA, 1969.

    [48] Schwantes E. The propulsion jet of a VTOL aircraft. In:

    AGARD conference proceedings CP-91, propulsion andenergetics panel symposium, Sandefjord, Norway, 1971.

    p. 15.110.

    [49] Ransom ECP, Smy JR. Introduction and review of some

    jet interference phenomena relevant to V/STOL aircraft.

    In: AGARD Report R-710, 1984. p. 2.123.

    [50] Kotansky DR. Jet flow-fields. In: AGARD Report

    R-710, 1984. p. 7.148.

    [51] Kotansky DR, Durando NA, Bristow DR, Saunders PW.

    Multi jet induced forces and moments on VTOL

    aircraft hovering in and out of ground effect. NADC

    Report 77-229-30, Naval Air Development Centre, June

    1977.

    [52] Ing DN, Zhang X. An experimental and computational

    investigation of ground effect lift loss for single jet

    impingement. Aeronaut J 1994;98(974):12736.

    [53] Wyatt LA. Static tests of ground effect on planforms

    fitted with a centrally-located round lifting jet. Aeronaut

    Research Council C.P. No. 749, Ministry of Aviation,

    1964.

    [54] Corsiglia VR, Wardell DA, Kuhn RE. Small-scale

    experiments in STOVL ground effects. In: Proceedings

    of the international powered lift conference, London,

    UK, 1990. p. III.14.113.

    [55] Welte D. Prediction of aerodynamic interference effects

    with jet-lift and fan-lift VTOL aircraft. In: AGARD

    conference proceedings CP-143, fluid dynamics panel

    symposium, Delft, Netherlands, 1974. p. 23.19.[56] Agarwal RK. Recent advances in prediction methods for

    jet-induced effects on V/STOL aircraft. In: Krothapalli

    A, Smith CA, editors. Recent advances in aerodynamics,

    New York : Springer; 1986. p. 471521.

    [57] Hancock GJ. A review of the aerodynamics of a jet in a

    cross-flow. Aeronaut J 1987;91:20113.

    [58] Margason RJ. Fifty years of jet in cross-flow research. In:

    AGARD conference proceedings CP-534, fluid dynamics

    panel symposium, Winchester, UK, 1993. p. 1.141.

    [59] Keffer JF, Baines WD. The round turbulent jet in a cross-

    wind. J Fluid Mech 1962;1(31):48196.

    [60] Shaw CS, Margason RJ. An experimental investigation of

    a highly underexpanded sonic jet ejecting from a flat plate

    into a subsonic cross-flow, Technical Note D-7314,

    NASA, 1973.

    [61] Stoy RL, Ben-Haim Y. Turbulent jets in a confined cross-

    flow. Trans ASME J Fluids Eng 1973; 5516.

    [62] Ransom ECP, Wood PM. A literature review on jets in

    cross-flow. In: AGARD conference proceedings CP-143,

    fluid dynamics panel symposium, Delft, Netherlands,1974. p. 26.17.

    [63] Taylor P, Watkins DJ. An investigation of inclined jets

    in a cross-wind. In: AGARD conference proceedings

    CP-308, fluid dynamics panel symposium, Fundac-ao

    Calouste Gulbenkian, Lisbon, Portugal, 1981. p. 6.14.

    [64] Bradbury LJS. Some aspects of jet dynamics and their

    implications for VTOL research. In: AGARD conference

    proceedings CP-308, fluid dynamics panel symposium,

    Fundac-ao Calouste Gulbenkian, Lisbon, Portugal, 1981.

    p. 1.126.

    [65] Kamotani Y, Greber I. Experiments on a turbulent jet in

    a cross-flow. AIAA J 1972;10(11):14259.

    [66] Chaissaing P, George J, Claria A, Sananes F. Physicalcharacteristics of subsonic jets in a cross-stream. J Fluid

    Mech 1974;62(1):4164.

    [67] Fearn RL, Weston RP. Induced velocity field of a jet in a

    cross-flow. Technical Paper 1087, NASA, 1978.

    [68] Ivanov Y. Shape of the centreline of an axisymmetric fan-

    type jet in a cross-flow. Izv Vuz Aviotsionnaya Teknika

    (4) (1963).

    [69] Shandorov G. Calculation of a jet axis in a drifting flow.

    Technical Translation F-10638, NASA, 1966.

    [70] Margason RJ. The path of a jet directed at large angles to

    a subsonic freestream. Technical Note D-4919, NASA,

    1968.

    [71] Zhang X, Hurst DW. An investigation of a supersonic jet

    in a low speed cross-flow. Report 301896, Department of

    Aeronautics and Astronautics, University of Southamp-

    ton, 1990.

    [72] Wooler PT. Flow of a circular jet into a cross-flow. AIAA

    J Aircraft 1969;6(3):2834.

    [73] Snel H. A method for the calculation of the flow-field

    induced by a jet exhausting perpendicularly into a cross-

    flow. In: AGARD conference proceedings CP-143, fluid

    dynamics panel symposium, Delft, Netherlands, 1974.

    p. 18.16.

    [74] Siclari MJ, Migdal D, Palcza JL. The development of

    theoretical models for jet-induced effects on V/STOL

    aircraft. In: Proceedings of the AIAA/SAE 11th pro-

    pulsion conference, Anaheim, CA, USA, 1975, AIAA75-1216.

    [75] Sucec J. Bowley WW. Prediction of the trajectory of a

    turbulent jet injecting into a cross-flowing stream. Trans

    ASME, J Fluids Eng (1976) 66773.

    [76] Adler D, Baron A. Prediction of a three-dimensional

    circular turbulent jet in a cross-flow. AIAA Journal

    1978;17(2):16874.

    [77] Fearn RL. Progress towards a model to describe jet/

    aerodynamic surface interference effects. In: Krothapalli

    A, Smith CA, editors. Recent advances in aerodynamics,

    New York : Springer; 1986. p. 40734.

    [78] Roth KR. Numerical simulation of a subsonic jet in a

    cross-flow. In: Proceedings of the international powered

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 189

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    16/17

    lift conference, Santa Clara, CA, USA, 1987. p. 42531,

    SAE 872343.

    [79] Baker AJ, Snyder PK, Orzechowski JA. Three-dimen-

    sional characterisation of the near-field of a VSTOL jet in

    a turbulent cross-flow. Comput Methods Appl Mech Eng

    1993; (102) 113.

    [80] Chiu SH, Roth KR, Margason RJ, Tso J. A numericalinvestigation of a subsonic jet in a crossflow. In: AGARD

    conference proceedings CP-534, fluid dynamics panel

    symposium, Winchester, UK, 1993. p. 22.114.

    [81] Chocinski D, Leblanc R, Hachemin J-V. Experimental/

    computational investigation of supersonic jet in subsonic

    compressible crossflow. In: 35th AIAA aerospace sciences

    meeting and exhibit, Reno, NV, USA, 1997, AIAA 97-

    0714.

    [82] Kennedy KD, Walker BJ, Mikkelsen CD. Comparisons

    of CFD calculations and measurements for a sonic jet in a

    supersonic crossflow. In: 36th AIAA/ASME/SAE/ASEE

    joint propulsion conference and exhibit, Huntsville, AL,

    USA, 2000, AIAA 2000-3313.[83] Margason RJ. Jet-induced effects in transition flight. In:

    NASA special proceedings SP-116, 1966. p. 17789.

    [84] Fearn RL, Weston RP. Induced pressure distribution of a

    jet in a crossflow, Technical Note D-7916, NASA, July

    1975.

    [85] Spreeman KP. Induced interference effects on jet and

    buried-fan VTOL configurations in transition. Technical

    Note D-731, NASA, 1960.

    [86] Volger RD. Interference effects of single and multiple

    round or slotted jets on a VTOL model in transition.

    Technical Note D-2380, NASA, 1964.

    [87] Volger RD, Goodson KW, Low speed aerodynamic

    characteristics of a fuselage model with various arrange-

    ments of elongated lift-jets. Technical Note D-7299,

    NASA, 1973.

    [88] Hammond AD, McLemore HC. Hot-gas ingestion

    and jet interference effects for jet V/STOL aircraft.

    In: AGARD conference proceedings CP-27, fluid

    dynamics panel symposium, Go ttingen, Germany, 1967.

    p. 8.127.

    [89] Schultz G, Viehweger G. Detailed experimental and

    theoretical analysis of the aerodynamic interference

    between lifting jets and the fuselage and wing. In:

    AGARD conference proceedings CP-150, fluid dynamics

    panel symposium, Rome, Italy, 1974. p. 24.113.

    [90] Nangia RK. Prediction & evaluation of vectored jets-

    induced interference on aircraft configurations in low-speed flight. In: Proceedings of the 19th congress of the

    international council of aeronautical sciences, Anaheim,

    CA, USA, 1994. p. 85970.

    [91] Winston MM, Weston RP, Mineck RE. Propulsion-

    induced interference effects on jet-lift VTOL aircraft. In:

    Proceedings of the AIAA/SAE 11th propulsion con-

    ference, Anaheim, CA, USA, 1975, AIAA 75-1215.

    [92] Margason RJ, Volger RD, Winston MM. Wind tunnel

    investigation at low speeds of a model of the Kestrel

    (XV-6A) vectored thrust V/STOL airplane, Technical

    Note D-6826, NASA, 1972.

    [93] Fu tterer H, Harms L. Jet interference measurements on a

    VTOL model with jet simulation by fans, In: AGARD

    conference proceedings CP-27, fluid dynamics panel

    symposium, Go ttingen, Germany, 1967. p 17.126.

    [94] Krenz G, Barche J. Jet influence on V/STOL aircraft

    in the transitional and high speed regime. In: AGARD

    conference proceedings CP-27, fluid dynamics panel

    symposium, Go ttingen, Germany, 1967. pp. 2.125.

    [95] Wiles WF. Jet-lift intakes. In: AGARDograph 103,specialists meeting, Tennessee, TN, USA, 1965.

    p. 55986.

    [96] Curtis P. The design and testing of short duct intakes for

    supersonic STOVL aircraft. In: Proceedings of the RAeS/

    IMechE conference on engine/airframe integration, Not-

    tingham, UK, 1992. p. 10.112.

    [97] Tape RF, Hartill WR, Curry LS, Jones TJ. Vectoring

    exhaust systems for STOL tactical aircraft. Trans ASME

    J Eng Power 1983;105(3):65462.

    [98] McMichael T, McKay K, Walker MJ, Fielding C,

    Lockley G, Curtis P, Probert B, Lee CS, Moretti G.

    Aerodynamic technologythe role of aerodynamic tech-

    nology in the design and development of modern combataircraft. Aeronaut J 1996;100(1000):41124.

    [99] Carbonaro M. Interference problems in V/STOL testing

    at low speeds. In: AGARD conference proceedings

    CP-174, fluid dynamics panel symposium, London, UK,

    1975. p. 40.121.

    [100] Koenig DG, V/STOL wind-tunnel testing. In: AGARD

    Report R-710, 1984. p. 9.171.

    [101] Harris AE. Test techniques for engine/airframe integra-

    tion. In: AGARD conference proceedings CP-498, fluid

    dynamics panel symposium, Fort Worth, TX, USA, 1991.

    1.117.

    [102] Winston MM. Induced interference effects on the

    aerodynamic characteristics of a 0.16-scale six-jet V/

    STOL model in transition. Technical Note D-5727,

    NASA, 1970.

    [103] Mineck RE, Margason RJ. Pressure distribution on a

    vectored thrust V/STOL fighter in the transition-

    speed range. Technical Memorandum X-2867, NASA,

    1973.

    [104] Mineck RE, Schwendemann MF. Aerodynamic charac-

    teristics of a vectored thrust V/STOL fighter in the

    transition-speed range. Technical Note D-7191, NASA,

    1973.

    [105] Haftmann B. Jet effects on forces and moments of a

    vstol fighter-type aircraft. In: AGARD conference

    proceedings CP-308, fluid dynamics panel symposium,

    Fundac-ao Calouste Gulbenkian, Lisbon, Portugal, 1981.p. 18.113.

    [106] Harris AE, Wilde GL, Smith VJ, Mundell ARG,

    Davidson DP. ASTOVL model engine simulators for

    wind tunnel research. In: AGARD conference proceed-

    ings CP-498, fluid dynamics panel symposium, Fort

    Worth, TX, USA, 1991. p. 15.110.

    [107] Saddington AJ. Mutual interference between jets

    and intakes in STOVL aircraft. EngD thesis,

    Cranfield University, RMCS, Shrivenham, UK, Septem-

    ber 1997.

    [108] Saddington AJ, Knowles K. Mutual interference

    between jets and intakes in STOVL aircraft. Aeronaut

    J 1999;103(1024):2816.

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191190

  • 8/10/2019 CFD A review of out-of-ground-effect propulsion-induced.pdf

    17/17

    [109] Saddington AJ, Knowles K. Jet/intake interference in

    short take off, vertical landing aircraft. AIAA J Aircraft

    2001;38(5):92431.

    [110] Hunt DL, Bruce RJ. The role of CFD in predicting

    vectored jet interference for STOVL aircraft. In: Proceed-

    ings of the international powered lift conference, London,

    UK, 1998. p. 11.112

    [111] Margason RJ, Correlation of STOVL jet induced effects with

    examples for configurations with supersonic convergent jets.

    In: Proceedings of the international powered lift conference,

    Williamsburg, VA, USA, 2002. AIAA 2002-5967.

    [112] Margason RJ, Gentry Jr GL. Aerodynamic character-

    istics of a five-jet VTOL configuration in the transition

    speed range. Technical Note D-4812, NASA, 1968.

    A.J. Saddington, K. Knowles / Progress in Aerospace Sciences 41 (2005) 175191 191