You are on page 1of 58

Progress in Aerospace Sciences 37 (2001) 419476

Modern helicopter rotor aerodynamics


A.T. Conlisk1
Department of Mechanical Engineering, The Ohio State University, Columbus, OH 43210-1107, USA

Abstract

The helicopter rotor wake is among the most complex uid dynamic structures being three dimensional and in many
cases unsteady. The wake begins at the blade(s) where the ow can be transonic near the blade tip and undergo
compressible dynamic stall. Farther down in the wake, the ow is essentially incompressible. Moreover, the rotor blades
undergo complex unsteady motions because of the necessity to balance moments; they are elastic as well. In this paper,
the fundamental aeromechanics of the wake and the ow on the blade is discussed and the primary methods of analysis,
computation, and experiment employed to uncover the physics of the rotor wake are described. r 2001 Elsevier Science
Ltd. All rights reserved.

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
2. Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
2.1. Fundamentals and performance regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
2.2. One- and two-dimensional models of the rotor wake . . . . . . . . . . . . . . . . . . . . 423
2.3. Rotor blades . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
2.4. Elements of two-dimensional airfoil aerodynamics . . . . . . . . . . . . . . . . . . . . . . 425
2.5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
3. Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
3.1. The NavierStokes equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
3.2. Boundary and initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4. Three-dimensional wake aeromechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
4.1. Representations of the rotor wake by the BiotSavart law . . . . . . . . . . . . . . . . . . 430
4.2. Grid-based wake calculation methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
4.3. Blade aeroelasticity and trim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
5. Experimental methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
5.2. Levels of the experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
5.3. Data types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
5.4. Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
6. Special topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
6.1. Bladevortex interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
6.2. Interactional aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465
7. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471

E-mail address: conlisk.1@osu.edu (A.T. Conlisk).


1
Partially supported by the NRTC Rotorcraft Center of Excellence at Georgia Institute of Technology and the US Army Research
Oce.

0376-0421/01/$ - see front matter r 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 3 7 6 - 0 4 2 1 ( 0 1 ) 0 0 0 1 1 - 2
420 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

1. Introduction locally important. In addition, as the blades rotate, the


tip vortex shed from one of the blades may collide with a
It is a common experience while driving on any following blade; this phenomenon is known as blade
expressway around the world that a helicopter may vortex interaction (BVI) and is a major source of rotor
often be seen overhead monitoring trac ow. Indeed, noise on the helicopter. BVIs are most severe in vertical
helicopters play a critical role in a wide variety of descent and landing. Interactions between a number of
activities from important military missions to emergency individual components of the helicopter also occur; two
medical transport. In this paper we examine the important interactions are main rotorfuselage interac-
fundamental uid dynamics and aeromechanics of the tion, and main-rotortail-rotor interactions. All of these
helicopter rotor wake which possesses a wide range of uid dynamic phenomena occur simultaneously with the
diverse uid mechanics problems from incompressible evolution and convection of a complex three-dimen-
inviscid ow to transonic and even supersonic viscous sional and, in general, unsteady wake ow.
ow. In this paper we focus on the complex uid dynamics
The complexity of the ow induced by a helicopter is of the rotor wake and the local ow on the blade. The
illustrated by the number of fundamental uid dynamic primary objective of this paper is to review the primary
research problems present and a summary sketch of methods by which the dominant features and structure
these problems is depicted in Fig. 1 [1]. First, at Mach of the rotor wake are identied. Other reviews of this
numbers typical of operation, the ow near the rapidly subject area discuss more specic features of rotor wakes
rotating blade is compressible while the ow in the wake and their computation [16]. Reichert [7] and Phillipe
of the helicopter rotor blades is substantially incom- et al. [8] have reviewed the European perspective on
pressible because of the rapid reduction in velocity. helicopter design. The two main helicopter-oriented
Indeed, the ow may be transonic or locally supersonic journals in which papers appear are the Journal of the
on the advancing blade side where the relative velocity American Helicopter Society and Vertica; many papers
past the blade is in the direction of the relative free have appeared in the AIAA Journal of Aircraft and
stream; thus near the tip of the blade shock waves can AIAA Journal and as seen in the reference list have been
form. On the retreating blade side where the relative presented at various AIAA conferences. Papers are
velocity past the blade is in the direction opposite to the also presented at the Annual Forum of the American
relative free stream, because of the necessity to balance Helicopter Society held every Spring and at the
moments (discussed in the next section), the angle of European Rotorcraft Forum held in the Fall.
attack must be larger than on the advancing side and the From a practical point of view it is necessary to omit
ow may be stalled; in this region, viscous eects are discussion of some of the uid dynamic problems

Fig. 1. A summary of specic ow problems which occur on a helicopter [1]. The helicopter is moving from left to right and the UN in
the gure is relative to the helicopter.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 421

depicted in Fig. 1; these include dynamic stall for which


several reviews are available including Carr [9,10] and
McCroskey [11]. In addition we will not discuss shock
boundary layer interactions. Moreover, while turbulence
is an issue in the calculation of rotor wakes, standard
turbulence closure models are usually used in this eld
and so we omit extensive discussion of this topic as well.
Finally, the aeroacoustics of the helicopter, which is a
critical design consideration, merits its own review
[1214].
The specic method employed in the calculation of the
rotor wake depends on the objective of the calculation.
Helicopter design is truly an interdisciplinary activity; as
will be seen the blades must be in constant motion in
order to balance the moments and so there is a uid
structure interaction problem. Aeroelasticity of the rotor
blades is also an important feature in the modeling of
the rotor wake. As is depicted in Fig. 1, there are regions
of supersonic ow on the blade during one portion of
the cycle and regions where the ow is stalled during
another part of the cycle. Downstream of the blade, the
ow is incompressible and vortical in character. Thus, in
developing a model of the wake, because there are a
wide variety of ow phenomena occurring simulta-
neously, there is always a tradeo between the accuracy
of the model and the ease of incorporating the results
into a comprehensive rotor design code; these codes
normally include the specication of blade motions for Fig. 2. A simplied description of a typical iteration scheme to
balancing moments and the elastic properties of the compute the wake of a single helicopter blade.
blades. A typical wake calculation procedure is depicted
in Fig. 2. Here it is noted that the blade motion and the
wake calculation are fully coupled.
Helicopter wake computations are most often accom- of the methodologies currently employed to elucidate
panied by comparison with experimental data. This the rotor wake structure. We discuss in some detail both
situation reects the truly coupled approach which takes vortex methods based on the BiotSavart law and grid-
advantage of insight gained from experiments to build based methods derived from the NavierStokes equa-
better theoretical and computational models of the tions or some approximation to these equations. We
rotor wake. For this reason, while experimental then discuss the aeroelastic properties of rotor blades
methods are normally not discussed in connection and typical trim equations. The types of experiments
with aerodynamics of the rotor wake, a section on that are commonly performed are presented next. We
experimental methods is included in this paper. It then discuss some related topics such as the BVI
should also be mentioned that often the knowledge ow problem and interactional aerodynamics. A summary
is reversed in the sense that physical phenomena follows.
uncovered in calculations suggest appropriate experi-
ments.
In a paper of this size it is impossible to avoid using 2. Overview
the same symbol for dierent variables. To avoid
confusion, care has been taken to dene each variable In this section, we introduce some of the basic
as it appears within a given section. Thus the meaning of terminology associated with rotorcraft and discuss some
a given symbol should be clear from the context of the basic principles of operation which aect the rotor wake
section. structure and are covered in various textbooks [1521].
The paper is organized as follows. In the next section We then discuss the origins of the analysis of the wake in
we discuss the fundamentals of operation of the one- and two-dimensional models, rotor blade shapes,
helicopter including a discussion of the need to balance followed by a discussion of those elements of two-
moments on the rotor blades. We then present the dimensional thin airfoil theory which are used in the
governing equations followed by a detailed presentation three-dimensional lifting-surface calculations.
422 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

2.1. Fundamentals and performance regimes

The helicopter can perform maneuvers such as vertical


takeo and landing and hover which are not possible for
xed-wing aircraft. The helicopter operates in four ight
regimes: hover in which the thrust generated by the rotor
blades just osets the weight; vertical climb and descent
and forward ight. Descent is a particularly dicult
regime because the rotor blades are pushed downward
into the vortical wake and considerable noise and
vibration may occur as a result. In forward ight the
component of the thrust in the forward ight direction
must overcome the drag. Forward ight is characterized
by the advance ratio, m V=OR where V is the forward
ight speed, O is the angular speed of the rotor and R is
the rotor radius. Typically design constraints suggest
mp0:4:
As with a xed-wing aircraft, a constant lift should be
produced during the steady ight of a helicopter.
However, since the rotor blades rotate in a single
direction, in forward ight, there will be a moment
imbalance. Consider the rotating motion of a single
helicopter rotor blade as depicted in Fig. 3. In this gure
c is the angle the blade makes with the horizontal axis
and is called the blade azimuth and x * and y * are inertial
Cartesian coordinates. As the rotor blade moves in the
same direction as the forward ight speed (the advancing
blade side), the velocity relative to the blade is higher
than the local rotational speed of the rotor and since the
lift is proportional to the velocity squared, the angle of
attack need not be large to achieve sucient lift. On the
other hand, as the blade moves in a direction opposite to
the direction of ight (the retreating blade side) the Fig. 3. A single rotor blade in forward ight; a top view.
velocity relative to the blade is smaller and the angle of (a) Advancing and retreating sides of the rotor disk. Here the
attack must thus be larger to achieve the same total lift. velocity V is the absolute velocity of the helicopter.
The process of balancing the forces and moments is (b) Denition of lag and ap angles. (c) Lift and drag in
called trim; that is, the angle of attack of the blades on forward ight; here a is the blade angle of attack measured from
the advancing and retreating sides must be adjusted the line of zero lift.
periodically throughout each blade rotation cycle so that
there is a balance of rolling moment; this is called cyclic quantities from which performance may be assessed. As
pitch and the rotor is said to be trimmed. The collective is well known, the total drag is composed of pressure or
pitch of the blades is a control in which the angle of form drag and viscous drag. In situations where loads are
attack of each of the blades is increased simultaneously generated by three-dimensional vortex systems, the
to achieve a higher lift; for a hovering rotor, an increase pressure drag is usually called induced drag. Lift is
of the collective pitch, for example, results in climb. comparatively easy to predict since it is usually found
In addition, rotor blades have a large span to chord from a surface pressure integration, although this is not
ratio and so, depending on how the blades are attached the case when the inuence of BVI is strong. On the
to the hub, large stresses can be communicated to the other hand, the power loss due to drag is very hard to
hub if the blades are not permitted to ap and bend. predict because it is a much smaller force and is thus
Indeed, blade aeroelastic eects play a major role in sensitive to small changes in pressure [22].
determining helicopter performance [20]. The major design parameters for the helicopter are the
The primary task in rotorcraft aerodynamic design is dimensionless coecients related to thrust and power.
to determine the lift and drag coecients of the rotor These coecients are the thrust coecient dened by
blades since these two quantities determine the thrust
and power required for given speed in forward ight or T
CT ; 1
hover. Thus the lift and drag coecients are the two rAOR2
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 423

the torque coecient dened by


Q
CQ 2
rAOR2 R
and the power coecient
P
CP ; 3
rAOR3
where A is the disk area, R is the rotor radius and P is
the power required to produce the thrust (T). Note that
since P OQ; CP CQ :
Hover performance is measured by value of the gure
of merit. This quantity is dened as the ratio of the ideal
power required to produce the thrust and the actual
total power required PA :
PI
FM ; 4
PA Fig. 4. Sketch of the rotor wake in descent based on a one-
dimensional momentum and control volume approach.
where the ideal power is dened as the actuator disk
result
simally thin control volume around the rotor disk, the
PI Tvi ; magnitude of the pressure drop across the disk is
where T is the thrust and vi the uniform downwash T
Dp 6
through the rotor disk. The actual power includes the A
prole power needed to overcome the aerodynamic drag and this pressure drop is called the disk loading.
of the blades and the induced power caused by the Now consider hover where V 0; then applying the
vortical wake. Typically a well-designed rotor can momentum and energy equations to a control volume
achieve FMB0:70.8. The diculty with determining which extends above the rotor disk to where the local air
the gure of merit is that the power required due to the velocity is zero and below the disk where the velocity is
presence of the wake (i.e. the induced power) is dicult vN we have
to calculate accurately.
In order to calculate the thrust and power coecients, Tvi 12m v2N m vN vi
some model for the rotor wake must be assumed. and it follows that vN 2vi : Moreover, applying
Current wake methodologies will be discussed in the Bernoullis equation above the disk, and then below
next section and in what follows we discuss simple one- the disk we have Dp 12rv2N : This indicates that the
and two-dimensional models for the rotor wake. rotor wake contracts as the uid velocity approaches vN
far from the rotor pdisk
and the wake p radius
far from
2.2. One- and two-dimensional models of the rotor wake the disk is rN 1= 2R; the factor 1= 2 is called the
contraction ratio. Moreover, applying Bernoullis equa-
The basic theory of the helicopter rotor was developed tion above the disk, and then below the disk we have
for the airplane propeller and subsequently applied to Dp 12rv2N T=A:
the helicopter. Consider the case of hover. The simplest The downwash vi can be calculated directly as
model for the rotor wake is to assume a one-dimensional r
p CT
ow through the rotor disk as depicted in Fig. 4; this is vi T 2rA OR
called momentum theory [2325] where the rotor is 2
considered to be an actuator disk generating a uniform or in dimensionless form
r
induced velocity vi at the rotor disk. If the ow is vi CT
assumed to be incompressible, inviscid and steady, then l :
OR 2
a simple force balance on a control volume enclosing the
rotor shows that the thrust is given by Thus, for hover,
3=2
T m vN  V; 5 C
CP CT l pT
where m rAvi rAN vN is the mass ow rate through 2
the rotor disk and V is the velocity of the rotor disk. The so that the gure of merit is given by [21]
quantity vi ; the velocity at the rotor disk is called the 3=2
C
induced velocity or downwash. The induced power in this FM p T ;
case is dened by P 12m v2N  V 2 : Using an innite- 2CP;meas
424 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

where CP;meas is a dimensionless measured power the section lift coecient, CL : In this case at any
coecient. Incorporation of experimental data in the blade section, where the local velocity is Or * ; dT
1 * 2
form of a correction factor k allows us to write 2rcCL Or dr where r is the dimensional spanwise
* *

3=2 location of the section and for N rotor blades


kC sCd0
CP;meas pT ; 1 Nb c 1
2 8 dCT CL r2 dr sCL r2 dr: 7
2 pR 2
where s Nb c=pR is the geometric solidity. The rst Here c is the blade chord and r r * =R: To obtain the
term in this equation is the dimensionless induced power thrust coecient Eq. (7) must be integrated along the
and the second term is the consumed prole power due span and
to section drag: Z 1
Z R Z 1
1 R CT s CL r2 dr: 8
P0 Dsec OR dr * rOr * 2 ccd ORNb dr * ; 2 0
0 2 0
For small angles of attack, particularly simple formulas
where c is the chord, or for CT can be deduced. Following a procedure similar to
P0 18rO3 R4 ccd0 Nb that described above, the result for the power coecient
is
assuming the drag coecient is constant at cd0 and the Z 1
blade is rectangular. In dimensionless form 1
CQ s lCL r2 CD r3 dr; 9
2 0
P0 1
CP0 scd0 : where l is termed the inow factor and for the case of
rAOR3 8
hover is given by l vi =OR; recall that vi is the induced
Despite attempts to improve momentum theory by velocity. In blade-element theory, the lift and drag
accounting for rotation in the wake [26] the deciencies coecients are obtained from lookup tables.
of the momentum approach are clear: momentum To compare performance of rotor blades with
theory assumes uniform ow in the rotor wake and variable wing planforms a thrust-weighted solidity is
contains no information about the specics of the rotor often dened by
blade. Moreover, the actuator disk model is only valid Z Z 1
as the number of rotor blades becomes large. 1 1 1
CT srCL r2 dr se CL r2 dr 10
The next step in a model of the rotor wake is to 2 0 2 0
incorporate the blade geometry into the problem and and assuming a constant lift coecient, and equal
this is called blade-element theory. The rst detailed thrust,
calculations of blade loading for the assessment of Z 1
performance were apparently performed by Drzewiecki se 3 sr dr: 11
(see [27], p. 179) who assumed that each slice of the 0
blade at a xed radial location behaves independently; A torque-weighted solidity may also be dened ([21],
while lacking a rm theoretical basis for this conjecture, p. 111).
he used the corresponding two-dimensional airfoil
properties to calculate performance. He found a 2.3. Rotor blades
signicant error in comparison with experiment because
he used only the average velocity upstream of the rotor Typical full-scale helicopter rotor blades have large
disk and neglected the velocity at the rotor disk induced span to chord ratios on the order of 1520. Model scale
by the rotation of the blade. Improvements in the rotor blades have much smaller aspect ratios for both
calculation of the induced velocity were made with the safety and practical reasons. Moreover, the planforms
advent of Prandtls Lifting Line Theory which was can be complex; Fig. 5 is a sketch of several rotor blade
under development at the time (see [27], p. 181). shapes. Fig. 5(a) is a UH-60 blade which is used on
The validity of blade-element theory was veried several helicopters. The tapered tip (Fig. 5(b)) is
experimentally [28] and put on a rm theoretical basis designed to minimize noise, while the same is true of
when it was shown that the velocity potential for the the British Engineered Rotor Planform (BERP) tip. In
potential ow relative to a rotating body at any cross addition, in an eort to achieve a uniform inow to the
section is equivalent to the two-dimensional non- rotor, rotor blades are twisted in the sense that the
rotating velocity potential [29]. This approach was then physical angle of attack varies with span. For a linear
extended to unsteady ow [30] and today, loads are twist
routinely calculated using two-dimensional section
a a0 atw y;
properties.
Provided the angle of attack (Fig. 3(b)) is below where a0 is the initial angle of attack, y measures
the stall angle, the thrust can be expressed in terms of dimensionless distance along the span and atw is
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 425

Fig. 6. Implementation of the two-dimensional Kutta condi-


tion on an airfoil when using surface vortex distribution. The
circulation per length at the trailing edge gT:E: which is the sum
of the corresponding upper and lower surface circulation per
length gU and gL ; must vanish. Also the trailing circulation per
length gW is aligned with the local ow direction.

Fig. 5. Several rotor blades: (a) UH-60; (b) rotor blade with a disturbance due to the bodys motion relative to the free
tapered tip; and (c) a British Engineered Rotor Planform stream ow must vanish far from the body.
(BERP). The equation r2 F 0 is satised in two dimensions
by source, doublet or vortex distributions [32,33]. The
no-normal ow boundary condition can be satised by
a constant. The ideal twist is the twist distribution that choosing the appropriate strengths of the singularity
gives constant inow in blade-element theory [20]. A distributions of source, vortex or doublet types.
popular cross section for rotor blades is the NACA0012 However, the solution for the velocity eld is not
airfoil whose maximum thickness is 12% chord. A unique even after selecting a desirable combination of
number of rotor blades and their properties are given singularity elements. According to a Helmholtz theorem,
in [31]. a vortex cannot start or end in the ow eld, so the
bound circulation on the wing must be shed into the ow
2.4. Elements of two-dimensional airfoil aerodynamics downstream. To make the solution unique, the wake
must be shed at the trailing edge to satisfy the Kutta
Since from blade-element theory the local ow at a condition (Fig. 6). One statement of the Kutta condition
given spanwise location is two dimensional, it is is that the ow is required to be parallel to the trailing
appropriate to discuss properties of the two-dimensional edge of the wing as it leaves, thus xing the circulation
ow past an airfoil. Most methods of solution use generated by the airfoil. Mathematically, the Kutta
singularity methods the basic principles of which are condition requires the vorticity component parallel to
used in three-dimensional lifting-surface rotor codes. In the trailing edge gT:E: to be zero:
what follows we present the analysis for the two- gT:E: gU gL 0; 13
dimensional ow past an airfoil immersed in a potential
ow. Although similar analysis has been published where gU and gL are the corresponding upper and lower
before (for example [32,33]), it is included here for surface vorticities expressed as circulation per unit
completeness. In this section, all variables will be chord. The Kutta condition also requires the shed
presented in dimensional form; the non-dimensionalized vorticity gW to be aligned with the local ow direction.
equations can be obtained by inspection. Fig. 6 illustrates the implementation of the two-dimen-
It is common knowledge that the two-dimensional sional Kutta condition for a vortex distribution on an
airfoil problem can be decomposed into a thickness airfoil.
problem and a lifting problem [32]. To illustrate the Consider the lifting portion of the problem, which for
calculation of a section lift coecient, we consider the simplicity we take to be a zero-thickness airfoil at an
linearized two-dimensional airfoil problem for the angle of attack a0 as shown in Fig. 7. The leading and
velocity potential (Fig. 6): trailing edges of the airfoil are located at x 0 and c;
Ftotal F FN 12 respectively and UN is the free stream velocity. While
sources or doublets can be used for the solution ([32],
2 2
and it follows that r F 0; where r is the Laplacian p. 119), it is common to consider the airfoil to consist of
in two dimensions. The total potential Ftotal must satisfy an unknown continuous vorticity distribution gx
two boundary conditions. First, the velocity normal to (circulation per unit length) on the airfoil. On the plane
the body surface vanishes: rFtotal  ~
n 0 where ~n is the z 0; the velocity induced at x0 by vorticity gx located
unit vector normal to the surface. Second, the ow at the point x is gx=2px0  x: Integrating gx from
426 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

equation with the assumptions (@F=@z and @F=


@x5UN and a0 51), we have
r
p  pN  u2  UN
2
17
2
E  rUN u  UN 18

@F
rUN ; 19
@x
where p is the local pressure and pN is the pressure at
innity. For a vorticity distribution gx; the streamwise
velocity in the linearized limit is given by
Z c
z
ux; z gx0 dx0 : 20
0 x  x0 2 z2
This integral can be evaluated using a change of variable
l x  x0 =z and we nd that @F=@xx; 0 gx=2
and @F=@xx; 0 gx=2: Hence the pressure dif-
Fig. 7. Continuous vortex distribution on a zero-thickness, ference between the upper and lower surface Dp is
symmetric airfoil. Note that the y-axis is normal to the xz Dpx pl  pu rUN gx: 21
plane and outward. G; the total circulation is the integral of g
over the streamwise direction. The lift L is obtained by integrating Dpx over the
airfoil:
Z c Z c
L Dpx dx rUN gx dx: 22
0 0
the leading edge (x 0) to the trailing edge (x c), the Substituting Glauerts [34] solution (15) into Eq. (22), we
linearized zero normal-velocity boundary condition at obtain the lift force acting on the airfoil:
the given point x0 becomes 2
Z L rUN pca0 23
1 c dx
gx UN a0 ; 14 which is recognized as the classical KuttaJoukowsky
2p 0 x0  x
law. The section lift coecient is dened by
where 0ox0 oc and a0 is the geometric angle of attack.
To render the solution unique, the two-dimensional L
cL 1 2 pa0 : 24
Kutta condition is applied, requiring gc 0: Eq. (14) is 2 rU Nc
a linear, singular integral equation for the unknown In hover, these equations can be directly applied to the
function gx: The analytical solution for Eq. (14) is rotorcraft problem by taking UN Or * at each radial
given by Glauert [34] as airfoil section. The dimensional velocity relative to the
1 cos y blades in forward ight is given by
gy 2UN a0 ; 15
sin y UN V sin c Or * ;
where y is dened by x c=21  cos y: For a nite-
where c is the azimuthal angle of the blade and V is the
thickness airfoil, the shape of the surface must be
velocity of the helicopter as in Fig. 3(a). Recall that the
accounted for and the appropriate equation is
Z   thrust coecient depends on lift coecient and so
1 c dx dz determination of the lift coecient is important.
gx UN a0  16
2p 0 x0  x dx For a cambered airfoil, there are additional terms in
and the boundary condition is satised along the mean Eq. (24), and many airfoil shapes are given as numerical
camber line. functions, and so the lift coecient is a numerical
Having obtained the distribution of vorticity on the function contained in lookup tables.
two-dimensional lifting airfoil, the pressure, lift and The section pitching moment M0 about the leading
moment can be calculated. The lift on the airfoil is edge is
Z c Z c
proportional to the pressure dierence between the
M0 Dpxx dx rUN gxx dx: 25
upper surface pu and lower surface pl : In a steady-state 0 0
ow, the velocity at any point in the ow eld is a
Hence for this zero-thickness airfoil
combination of the free stream velocity and the
2
perturbation velocity, as u; v UN cos a0 @F=@x; 2 c
M0 rUN p a0 : 26
UN sin a0 @F=@z). Substituting into the Bernoulli 4
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 427

Fig. 8. The lumped-vortex element structure for a single at-


plate airfoil at angle-of-attack.
Fig. 9. Point vortex representation of the thin, lifting airfoil.

Similarly, the moment coecient is dened by


element satises the Kutta condition automatically, the
M0 1 last panel (panel 5 in the gure) inherently fullls this
cm 1 2 2
 pa0 : 27
2rUN c
2 requirement and no additional specication of the Kutta
The center of pressure xcp ; where the moment is zero condition is needed. We write the solution for the
(this also can be considered to be the point where the lift velocity eld as the sum of singularities located at the
force acts), can be obtained by balancing the pitching center of pressure; in particular at small angle of attack,
moment and the product of the lift and is given by the velocity normal to the airfoil at the collocation point
xk may be written as
M0
xcp : 28 X
N
L
Wk Ajk Gj ; k 1; y; N; 30
Substituting M0 and L into Eq. (28), we nd that for a j1
zero-thickness airfoil at an angle of attack a0 ; xcp 14c;
that is, the center of pressure xcp is located at the one- where Gj is the circulation of the singularity on panel j
quarter chord. Because the lift acts at the one-quarter and Gk gk Dxk : The normal velocity induced by panel j
a single point vortex with the strength G is dened as the inuence coecient Ak; j ; j 1; y; N:
Rchord,
c The normal velocity at each collocation point must
0 gx dx will give the correct lift and can be used to
replace the continuous vortex distribution gx: Since the vanish and so
vorticity distribution gx is represented by a single point X
N

vortex of strength G; the zero normal-velocity boundary Ajk Gj wN;k a0 : 31


j1
condition at the surface can be specied correspondingly
at a single point, termed a collocation point and the Here wN;k is the vertical component of the free stream
location of this point is obtained by solving velocity and is constant for a xed wing. The inuence
G coecients are given by
UN a0 29
2pxcol  xcp xj  xk
Ajk ;
and xcol 34c as shown in Fig. 8. In Fig. 8, the xj  xk 2 z2k
concentrated point vortex is located at the one-quarter where zk is the location of the mean camber line at xk :
chord and the collocation point at the three-quarter The values of Gj ; j 1; y; N can be computed numeri-
chord and this two-dimensional point-vortex singularity cally by solving this fully dense-matrix equation. Having
may be called a lumped vortex element. It should be solved for the value of Gj ; when assuming incompres-
noted that this two-dimensional analysis forms the basis sible ow, the pressure on the surface and the uid
for three-dimensional lifting surface models of a rotor dynamic loads can be calculated using Bernoullis
blade. equation.
In the general case where the airfoil section may be The other performance parameter which is required to
represented by a numerical function, the solution for the calculate the power coecient is the drag coecient
velocity eld must be calculated numerically. It is dened by
customary to divide a thin airfoil into N panels and
D
each panel has been replaced by a lumped vortex cD 1 2 ; 32
element as shown Fig. 9. Because the lumped vortex 2rU Nc
428 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

where D is the drag. The drag has two components, the is often poor. In what follows, we present the modern
viscous drag and the drag induced by the trailing vortex three-dimensional methods which are commonly used to
system, called the induced drag. For an airfoil at low calculate the velocity eld induced by a rotor wake.
angle of attack, such as, may occur on the advancing
side of the rotor, these two components of the drag may
be comparable. On the other hand, on the retreating side 3. Governing equations
where the ow may be massively separated, the induced
drag may be far larger than the viscous drag. Note that 3.1. The NavierStokes equations
the induced drag can be calculated assuming inviscid
ow, while the viscous drag or skin friction drag must be In discussing modern three-dimensional approaches
calculated from the solution of the viscous boundary to modeling the rotor wake, it is instructive to begin with
layer equations on the blade. Typically, these solutions the statement of the governing equations of compres-
for various blade shapes are catalogued and the viscous sible uid ow which are the continuity equation along
drag coecients are available in lookup tables. with the NavierStokes equations and the energy
Compressibility can be incorporated into the lift, drag equation. In a xed, inertial coordinate system these
and moment coecients by dividing the incompressible equations in tensor form are
result by the PrandtlGlauert compressibility factor b @r @ruj
dened as 0; 34
@t @xj
p
b 1  Ma2 ; 33 @rui @rui uj @p @tij
 ; 35
@t @xj @xi @xj
where Ma is the relative Mach number [20], pp. 262263,
the ratio of the relative velocity to the speed of sound,
p @rh @rhuj @p @p @ui @qj
a kRT ; where k is the ratio of specic heats, R is the  uj tij  ; 36
@t @xj @t @xj @xj @xj
gas constant and T is the local temperature.
where r is the density, p is the pressure, h the specic
enthalpy, tij the stress tensor, uj the velocity eld and qj
2.5. Summary
the heat ux. Here the repeating index indicates a
summation and both i and j are in the range 1 to 3. The
As performance requirements increase, more complex
dependent variables in these equations are the velocity
models of the rotor wake are required to calculate the
eld, the pressure, density, and the heat ux which is
induced loads. A barrier to the accurate and ecient
dened as
calculation of the rotor wake is the presence of a wide
range of important length scales (and hence time scales) @T
qj k ;
from the blade length to the size of the trailing vortex @xj
core; these length scales can span several orders of where k is the thermal conductivity. Assuming air is a
magnitude. Thus modeling and experimentation of Newtonian uid and a perfect gas, the stress tensor is
helicopter ows are extremely challenging, time con- related to the velocity eld by
suming and costly. Because of these complexities it is  
dicult to incorporate the dynamic nature of the entire @um @ui @uj
tij ldij m ; 37
rotor ow in the presence of the helicopter airframe @xm @xj @xi
in one single numerical computation or experimental where dij 0 for iaj and dij 1 for i j is the
measurement program. For this reason, rotorcraft Kronecker d and l 23m where m is the dynamic
research tends to be focused on one or two specic viscosity.
aspects of the rotor ow eld and tends to have both To complete the system, for a perfect gas,
experimental and computational components. For
example, many computational and experimental ap- p rRT 38
proaches have focused on the rotor wake ow for the and
case of two or four rigid blades rotating at relatively low
h cp T; 39
tip speeds. Under these conditions, it is often not
dicult to obtain good results for the blade pressure where cp is the constant-pressure specic heat.
distribution. On the other hand, at high tip speeds under The above system of eight equations in eight
forward ight and descent conditions this is often not unknowns along with the associated boundary and
possible and much more fundamental understanding of initial conditions are almost never solved in the form
these ight conditions is required. Moreover, numerical given above because of their complexity; indeed, blade
calculations may compare well with model experiments and wake computations are dierentiated by the actual
whereas under actual ight conditions, the comparison equation set solved and the choice of equation set
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 429

depends on the region of interest. If the origin of the 2 3


ru
wake is of interest, the ow on the blade needs to be 6 2 7
6 ru p  txx 7
calculated; this is usually done by using a grid-based 6 7
approach. The wake is modeled by some simple means. F6
6 ruv  txy 7;
7 42
6 7
A deciency of this approach is that due to numerical 4 ruw  txz 5
diusion, the vortical character of the wake is lost; the E pu  utxx  vtxy  wtxz qx
tip vortex, for example, most often cannot be identied
2 3
after about one rotor radius below the disk. On the other rv
hand, if the bulk of the wake is of interest, then a vortex 6 7
6 ruv  txy 7
approach, which preserves the vortical character of the 6 7
wake is appropriate and the ow on the blade is G6
6 rv2 p  tyy 7;
7 43
6 7
modeled. Both these approaches are discussed in the 4 ruw  tyz 5
sections to follow. E pv  utxy  vtyy  wtyz qy
Eqs. (34)(36) are often solved using a grid-based
2 3
method in a coordinate system rotating with the blade. rw
For a constant angular speed, the velocity relative to the 6 7
6 ruw  txz 7
blade in the moving coordinate system is given by 6 7
H6
6 ruw  tyz 7;
7 44
U0 U  X r; 6 7
4 rw2 p  tzz 5
E pw  utxz  vtyz  wtzz qz
where the bold type indicates a vector and the
coordinate systems are related as where the velocity vector is u; v; w; qx q1 in tensor
notation and t is the usual Newtonian viscous stress
X0 U0 t RtX; t0 t; tensor. Row 1 of each of the vectors dened above
corresponds to the continuity equation with the next
where the prime denotes the moving coordinate system
three rows corresponding to the momentum equations;
and U0 is the speed of advance; X x; y; z with a
the last equation corresponds to the thermal energy
similar denition of X0 : Here Rt is the rotation matrix
equation. Here E re V 2 =2 PE is the total energy
2 3 per unit volume, and e is the internal energy per unit
cos Ot sin Ot 0
6 7 mass. The symbol PE denotes potential energy and V is
4 sin Ot cos Ot 0 5:
the velocity magnitude.
0 0 1 In helicopter aerodynamics, the NavierStokes equa-
tions are often written splitting the inviscid and viscous
Note that Ot denes the angle c in Fig. 3(a) which
terms as
measures how far the blade tip has moved in a given
time t in the azimuthal direction; this distance is @U @F @G @H @Fv @Gv @Hv
45
@t @x @y @z @x @y @z
s Rc;
where Fv ; Gv ; Hv are the viscous stresses. This notation is
where R is the rotor radius. convenient for writing the thin-layer NavierStokes
In practice, the NavierStokes equations are not equations for which only the derivative normal to a
solved in the tensor form given above. Because a matrix solid surface is retained. If a solid surface is dened as
of equations will eventually be solved, it is convenient to z 0 then the thin-layer NavierStokes equations are
write the NavierStokes equations in vector form and given by
in most cases, they are solved in strong conservation @U @F @G @H @S
; 46
form as @t @x @y @z @z
@U @F @G @H where S is the vector
0; 40 2 3
@t @x @y @z 0
6 @u 7
where 6 m 7
6 @z 7
2 3 6 7
r 6 @v 7
6 m 7
6 7 S6 6 @z
7:
7 47
6 ru 7 6 7
6 7 6 4 @w 7
U6 7
6 rv 7; 41 6 m 7
6 7 6 3 @z 7
4 rw 5 4 5
m @ 2
E u v2 w2 ?
2 @z
430 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

The compressible NavierStokes equations, even in


the thin-layer approximation are extremely dicult to
solve and few grid-independent solutions of the full
equations for a rotor exist. Thus various approximations
are made to facilitate solutions. These approximations
are summarized in Fig. 10 going from the most
complicated physical situation to the least complicated.
Wake models based on each of the sets of equations
have appeared in the open literature and the details of
each method are described in the next section.

3.2. Boundary and initial conditions

The boundary conditions associated with the ow due


to a rotor pose complications not present in xed-wing
ows. First, far from the rotor disk in each direction, the
rotor wake draws uid from above the rotor disk and so Fig. 10. Summary of the various approximations used to
an inow boundary condition is required. This issue simplify the NavierStokes equations.
will be discussed later in the paper in the section on
grid-based methods. Second, the rotor blades undergo
complicated motions corresponding to a time- and
4. Three-dimensional wake aeromechanics
space-varying angle of attack due to trim requirements
and twist respectively in addition to rigid apping and
There are two primary approaches toward elucidating
leadlag motions in which the rotation speed of the
the fundamental features of the rotor wake theoretically
rotor varies with time and span. Moreover, rotor blades
and computationally. These approaches are (1) vortex
are aeroelastic and so may bend from root to span.
methods in which the wake is viewed as being inviscid
Consequently, both the no-normal ow and no-slip
but possessing vorticity which resides in the inboard
conditions are complicated since the velocity of the
sheet, the tip vortices and root vortices (Fig. 11) and (2)
blade is a complicated function of space and time. If
grid-based methods based on nite-dierence ap-
x denotes the chordwise direction, y spanwise and
proaches in solving the NavierStokes, Euler equations
z vertical and the rotor blade surface is dened by
or the full potential equation. In what follows we discuss
both of these approaches. The vast majority of the
z Sx; y; t:
calculations employed in the design of rotorcraft are
performed with an isolated rotor blade and in this
Then the unit normal to the rotor blade surface is section, interactions of the wake with other portions of
dened by the helicopter are not considered; those interactions are
considered in Section 6.2.
rF
n# ; As mentioned in the introduction, it is important to
jrFj note the notational diculties associated with discussing
both of these approaches in a single paper. In what
where F z  Sx; y; t: Thus the no-normal ow follows we use notational conventions appropriate to
boundary condition corresponds to each area. For example, in the vortex method work,
vectors are denoted by an over-arrow. In grid-based
u . n# U
~ ~blade xs ; ys ; zs ; t . n# 0:
methods, vectors are denoted by bold-face type. It is
hoped that the meaning of each of the variables is clear
For viscous ow, the tangential velocities must also within the context of its use.
equal the blade velocity. For inviscid ow, only the
normal ow condition can be imposed and implementa-
tion of the boundary conditions on the blade also 4.1. Representations of the rotor wake by the
depends on the nature of the local ow eld. Additional BiotSavart law
boundary conditions on the pressure can be obtained by
evaluating the governing equations at the boundaries of 4.1.1. Background
the domain. Far below the rotor disk, an outow As noted previously, actuator disk theory assumes
boundary condition is required. Finally, for unsteady that all of the dependent variables such as the velocity
calculations, an initial condition is required. and the pressure are uniform across the rotor disk.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 431

constant, the vortex system in the rotor wake can be


represented by the incompressible BiotSavart law.
Recall that for incompressible ow, the velocity eld is
divergence free which means that
r.~
u 0; 48

where ~ u is the velocity. From the denition of the


vorticity eld as the curl of the velocity, there exists a
~ [40] dened by
vector potential B
~ ~
ur B 49

and hence
~ rr . B  r2 B
r rB ~ r2 B
~ o
~; 50

where o ~ r ~
~ is the vorticity, o u: This is the three-
dimensional Laplace equation and in the absence of any
boundaries, the solution is
Z
~ 1 ~0 dV 0
o
Fig. 11. Sketch of a helicopter rotor wake for a single blade as B ; 51
envisioned by Gray [38,39]. 4p jX ~ X ~0 j

where the integral is over the volumetric region where


Clearly this is not the case and the inuence of the the vorticity is non-zero. Eq. (51) is written in Cartesian
number of blades was identied for the case of a coordinates and the velocity eld is obtained from the
propeller ([27], p. 181; also [35]) long ago. A simple denition of B ~ according to
model replaced the vortex sheet shed from each blade by !
Z 0
a series of straight parallel lines a xed distance apart ~V r B ~ 1 ~
o
U rX dV 0 : 52
leading to the denition of a tip loss factor ([27], p. 265) 4p ~X
jX ~0 j
which is still in use today. Goldstein [35] followed with
his description of the propeller wake as a series of Performing the dierentiation, in the integral we obtain
helicoidal surfaces of constant helix angle. These Z ~ ~0
surfaces were assumed to move downward at an average ~V  1 X  X o ~0
U dV 0 53
velocity obtained from momentum theory; the Goldstein 4p ~X
jX ~0 j3
analysis was then rened by allowing for a non-uniform
downwash [36]. This work forms the basis for the which is the standard form of the BiotSavart law. This
development of methods to predict performance char- formula forms the basis for all vortex calculations and it
acteristics of helicopter systems based on the existence of is the shape of the region over which the vorticity is non-
the vortex wake [37]. zero which determines what type of integral must be
evaluated. Because the BiotSavart law is only valid for
incompressible ow, rotor codes generally employ the
4.1.2. The BiotSavart law
PrandtlGlauert compressibility correction as described
It has long been known that the rotor wake is
in Section 2.4.
composed of high-strength tip vortices which roll up
downstream of each wing tip, an inboard vortex sheet
and a weaker root vortex as depicted in Fig. 11 [38,39]. 4.1.3. The tip vortex
The vorticity in the inboard sheet and the tip vortex is The tip vortex is by far the dominant inuence in the
conned to very thin regions which are surrounded by evolution of the rotor wake. Eq. (53) is a volume integral
substantially irrotational uid. This makes experiments and thus would be dicult and expensive to evaluate in
as well as computations extremely dicult because of its present form. Consider, rst the tip vortex. The tip
the rapid variation in velocity near the inboard vortex vortex is formed as a result of the rapid change of the
sheet, the tip vortices and near the airframe. The sense of bound circulation from a nite value to zero near the tip
circulation of the inboard sheet is opposite to that of the of the blade and is usually assumed to be a line vortex
tip vortex and has magnitude dGB =dy * where GB is the with vorticity distribution [40]
dimensional bound circulation on the blade and y * ox; y; z Gdx  x0 dy  y0 dz  z0 ;
measures distance outward along the span.
Assuming that the region outside the vortex sheet where x0 ; y0 ; z0 is the space curve which denes the
and the tip vortex is potential and the temperature is vortex position, and d is the Dirac delta function.
432 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Substituting into the BiotSavart law we obtain


Z ~X ~0 @X~0 0
~  1
~V X X
U G ds ; 54
4p C jX ~X ~0 j3 @s0
0 0
where X ~ X ~ s x0 ; y0 ; z0 and s is the parametric
variable dening the line vortex along curve C: Note
that the circulation of the tip vortex has been left inside
the integral to allow for its variation; in forward ight,
the circulation of the tip vortex varies with azimuth; in
hover the tip vortex circulation is a constant. Eq. (54) is
a line integral which can be evaluated numerically by a
standard integration procedure such as the Trapezoidal
rule or Simpsons rule. Alternatively, the simple formula
to be discussed in Section 4.1.7 may be used. However, Fig. 12. The inboard sheet as a collection of vortex boxes
expression (54) is singular at one point for each value of [44] modeled by straight line segments.
X~: This is a manifestation of the fact that the inuence
of viscosity in the core cannot be neglected [41] and this
leads to the use of the modied BiotSavart law [42,43]: depicted in Fig. 12 is
Z N X
X M

~V s; t  1
~ X
X ~0 @X ~0 =@s0 ds0 ~  1
~V X
U wei; j f~i; j Dxi Dyj ; 58
U G ; 55 4p
4p C fjX ~X ~0 j2 m2 g3=2 i1 j1

where m is a cuto parameter. Since any vortex may where f~ is the integrand of Eq. (57) and wei; j is the
locally be approximated by a parabolic arc with a well- weight function. Dierent weight functions correspond
dened radius of curvature, the value of the cut-o to dierent methods of integration. The simplest scheme
parameter is usually obtained by matching the local is to take wei; j 1 which assumes that the integrand is
viscous ow to that for a vortex ring [42,43]: piecewise constant over each segment. Note that the
m Z Z discretization can be interpreted as a collection of
1 4p2 a * 02 8p2 a * 02 rectangular-shaped vortex rings; this idea has been used
ln   2 v r dr 2 w r dr; 56
a* 2 G 0 G 0 by Egolf [44] who used the term vortex box to
where v0 and w0 are the velocities in the core of the describe these rings and calculated performance in
vortex. The value of m is subject to a model for the core forward ight. Often the shed vorticity component is
ow. In most rotor analyses, the inuence of the axial neglected (Fig. 13) leading to the inboard sheet being a
velocity w0 is neglected. Some vortex core models collection of line vortices with circulation gi; j Dxi :
commonly used are discussed in Section 4.1.5. A more accurate approximation to the inboard sheet
is to use curved vortex elements [45]. This situation is
4.1.4. The inboard vortex sheet depicted in Fig. 14 where again the shed vorticity
The simplest model for the inboard sheet is a component is neglected. The advantages of using curved
collection of straight line segments as depicted in vortex segments and an evaluation of the straight line
Fig. 12. From the BiotSavart law, if the vorticity segment approximation is given in Bliss et al. [45]. The
distribution is conned to at most a sheet with area A; decision of which of these two discretization schemes to
then the velocity distribution for the inboard sheet is use often depends on the aordability of computation
given by [40] time, especially if the wake module is imbedded within a
comprehensive rotor design code.
Z ~ ~0
~  1
~V X X  X ~g0 ~0 ;
U 03 dAX 57
4p A jX ~X ~j 4.1.5. Vortex core models
Most analyses of the rotor wake which use vortex
where ~
g is the strength density of the sheet, methods require a model for the azimuthal velocity in
Z
the vortex core. A number of models have been used and
g o
~ ~ dn;
several have been discussed by Leishman et al. [46]. The
most common is perhaps the simple Rankine vortex for
where n is the direction normal to the sheet.
which the core swirl velocity is
Because of the complexity of the integrand, the 8
integral in Eq. (57) must usually be computed numeri- > G r
>
< for rpav ;
cally. Consider a simplied two-dimensional integral 0 2p a2v
v 59
consisting of a rectangular sheet. Then the induced > G
>
: for rXav :
velocity due to the collection of vortex segments 2pr
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 433

where n is an integer. Leishman et al. [46] suggest that


the Vatistas vortex with n 2 t their experimental data
the best.
As with xed-wing vortices, the vortical nature of the
wake can persist for a signicant distance below the
rotor disk. Consider a Lamb-type velocity prole as a
model for the tip vortex:

G 2
uy 1  er =4ntt0 ; 63
2pr

where n is the kinematic viscosity, and the vortex was


created at time t0 : Then the point of maximum velocity
which is nominally considered
p to be the vortex core
Fig. 13. The inboard sheet as a collection of line vortices radius is given by av B 5nta where ta is the age of the
modeled by straight line segments with the shed wake neglected. vortex. Consider the vortex core radius to be about
the thickness of a rotor blade av B0:02 m: Then at a
rotational speed of about 250 rpm; the time it takes for
the vortex to diuse to twice its radius is given by

t2;a 4  1ta 3 5:3 s 15:9 s:

At a rotational speed of about 250 rpm; it would take


O t2;a B66 revolutions below the blade for the vortex
to diuse to twice its core radius. This number will vary
with the particular vortex core model chosen and may be
altered somewhat by the presence of turbulence;
however, since only a few revolutions of the rotor
(35) are of interest in most computations, it is clear that
for computational purposes, the tip vortex will not
diuse a signicant amount.

Fig. 14. The inboard sheet as a collection of line vortices


4.1.6. The time-dependent motion of an arbitrary
modeled by curved line segments.
three-dimensional line vortex
The crucial dierence between the motion of a two-
dimensional straight vortex lament (i.e. a line vortex)
Another distribution was used by Scully [47] in which and a three-dimensional lament is the presence of a
the swirl velocity is given by local self-induced velocity eld due to the nite
G r curvature of the vortex. In three dimensions a bent
v0 for rpav : 60
2p r2 a2v vortex will move even in free space whereas in two
dimensions a straight vortex will not.
In Eqs. (59) and (60), r is the local radial coordinate
The position of any material point on the vortex may
measured from the center of the vortex and av is the
be written in Cartesian coordinates as
vortex core radius. Little dierence is observed in the
proles when compared with experimental data ~0 x0 s; ti# y0 s; tj# z0 s; tk#;
X 64
although the Rankine prole seems to overpredict the
maximum velocity in the vortex when compared with
where t is the time. From the BiotSavart law, Eq. (54),
data [46]. Another distribution which has been used in
the three components of the vortex-induced velocity eld
rotorcraft calculations is the LambOseen vortex for
are, in general form,
which the swirl velocity is given by
G 2 2 UVx s; t
v0 1  e1:256r =ac : 61  
2pr @y0 @z0
0 0
Z z  z  y  y ds0
Finally, Vatistas et al. [48] used the distribution G @s0 @s0
;
G r 4p C fx  x0 2 y  y0 2 z  z0 2 m2 g3=2
v0 1=n
; 62
2p r2n a2n
v 65
434 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

UVy s; t
 0 
@z 0 @x0 0
Z x  x  z  z ds0
G @s0 @s0
;
4p C fx  x0 2 y  y0 2 z  z0 2 m2 g3=2
66

UVz s; t
 0 
@x 0 @y0 0
Z y  y  x  x ds0
G @s0 @s0
:
4p C fx  x0 2 y  y0 2 z  z0 2 m2 g3=2
67
In a Lagrangian time-marching approach to the Fig. 15. Vortex line segment geometry for calculation of the
calculation of the rotor wake the tip vortex would be three-dimensional velocity eld induced by the segment.
discretized and each point on the vortex advanced
forward in time according to the equation
~
@X The ends of the vortex line segment points are denoted
~;
s; t U 68
@t by pa and pb : In this section we calculate the three-
~ is the total velocity eld given by dimensional velocity eld induced by this nite line
where U
segment in the method rst described by Scully [47]. The
~U
U ~M U
~V U
~B ; 69 velocity distribution so calculated may be used to model
~M denotes any mean streaming motion and U ~B the bound circulation on the rotor blades and so the
where U
circulation is allowed to vary linearly from Ga at point pa
is due to the blades, any other bodies in the wake, and
and Gb at point pb :
that due to a far-eld boundary condition. In general,
The induced velocity is computed by using the Biot
the initial position of the tip vortex can be specied by
Savart law
X~s; t 0 f si# gsj# hsk#: 70
G~
ds ~
r1 Gsin y ds
Note that Eq. (68) is an integro-dierential equation dUV 3
; 71
4pr1 4pr21
because the velocity U ~V is given by the BiotSavart law.
Given a discretization of the spatial integral in the Biot where dUV is the induced velocity at point p and is
Savart law (a straight line approximation, Trapezoidal perpendicular to the rds plane, ds is an element of the
rule, Simpsons rule, etc.), we could then use an Euler or vortex line, G is the circulation at ds; r is the distance
modied Euler or the more accurate RungeKutta and from point p to ds; and y is the angle between r and the
AdamsMoulton routines to advance the vortex. Often direction of the vortex line segment (Fig. 15).
because the aerodynamic calculation is but a small part Further, dene h as the perpendicular distance from
of the overall rotorcraft design problem, the less point p to the vortex line segment and f as the angle
accurate but computationally less demanding Euler or between h and r (Fig. 15). Then,
modied Euler schemes are used. h
Complicating the computational issue, it is known sin y cos f ; 72
r
that a single helical vortex is linearly unstable to certain  
physical perturbations in hover and low-speed forward h
ds dh tan f df: 73
ight [4951]. Moreover, large-scale physical irregula- cos2 f
rities in wake vortex motions have been demonstrated Now using relationship (72), Eq. (71) can be rewritten in
experimentally [45,5254]. These physical irregularities terms of h and f:
take the form of interactions between tip vortices shed
from dierent blades and numerical instabilities if the G
dUV cos f df: 74
time step is not small enough. 4ph
In general, the circulation G can vary along the vortex
4.1.7. The velocity induced by a straight, nite vortex and we assume it can vary linearly from Ga at point pa to
line segment Gb at point pb ; so that
Many rotor codes model the tip vortex as a collection
G A B tan f;
of straight line segments. This method has the advantage
of being computationally ecient. Consider a straight where A and B are constants. Since tan f s=h we have
vortex line segment representing a rotor blade in Fig. 15. a linear variation in s along the vortex. Determining the
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 435

constants A and B we have


  cot a aj2  ~
j~ a  b~
cot a tan f 2 ; 89
G Ga Gb  Ga : 75 aj jb~j  2~
cot a cot b j~ 2
a  b~
cot a cot b
Substituting in Eq. (74) and integrating from sin a  sin b a b~j
j~ jb~j  j~
aj
: 90
f a  p=2 to p=2  b hcot a cot b j~ ~
aj jbj j~ 2 ~ 2
aj jbj  2~ a  b~
Z   
Ga p=2b Gb  Ga Substituting Eqs. (88)(90) into Eq. (79), and substitut-
UV cos f
4ph ap=2 Ga ing the result into Eq. (78), nally we haver
  (" #
cot a cos f sin f
df 76 ~ Ga j~ aj jb~j j~aj jb~j  ~ a  b~
cot a cot b UV
4p aj jb~j j~
j~ a b~j2
"   !#
or
   Gb  Ga aj2  ~
j~ a  b~
Ga Gb  Ga 1
UV cos a cos b Ga aj2 jb~j2  2~
j~ a  b~
4ph Ga   " #)
  Gb  Ga jb~j  j~
aj
sin a  sin b cot acos a cos b a b~:
~
: 77 Ga aj jb~jj~
j~ aj2 jb~j2  2~ a  b~
cot a cot b
The velocity UV is in a direction perpendicular to the 91
rds plane and hence perpendicular to the plane of the Consider the case where Gb Ga ; a semi-innite line
a and b~: Now dene positive circulation G such
vectors ~ vortex is obtained for say b 0 and the result is
~
that U V points into the paper as shown in Fig. 15. Now
~V can be written as G
U UV 1 cos a
4ph
~V UV ~ a b~ or
U 78
a b~j
j~ !
G aj2  ~
j~ a  b~
and thus UV 1 :
    4ph aj jb~ ~
j~ aj
Ga cos a cos b Gb  Ga cot a
UV 1 An innite line vortex is obtained in the limit as both a
4p h Ga cot a cot b
   and b approach zero and this result is
Gb  Ga sin a  sin b
: 79 G
Ga hcot a cot b UV :
2ph
To obtain the three-dimensional induced velocity U ~V
in terms of ~ ~
a and b; dene ~ c (Fig. 15) as
4.1.8. Relaxation techniques
~ ~
c b~ a: 80 Time-marching methods can be extremely accurate if
the time step is small enough. However, because the
Using the dot and cross-product relations and Fig. 15
wake module is only a small portion of the total design
~
a ~
c j~
aj j~
cj cos a; 81 process, rotor codes are often limited to time steps as
large as 15301 of motion of the blade. At these large
b~ ~
c jb~j j~
cj cos b; 82 time steps, time-marching methods are susceptible to
large numerical error. The alternative is to use an
j~
a ~
cj j~
aj j~
cj sin a; 83
iterative procedure in which spatial periodicity is
jb~ ~
cj jb~j j~
cj sin b; 84 enforced as a boundary condition to relax to the
steady freewake solution, provided such a solution exists
j~
a ~ cj [5559]. The idea behind relaxation techniques is to
h j~
aj sin a 85 write the position vector of the vortex as a function of
j~
cj
two variables
and from the denition of ~ c we have
~X
X ~a; t
j~
a ~ a b~ ~
cj j~ a b~j;
aj j~ 86
and so [57]
jb~ ~
cj jb~ b~ ~ a b~j:
aj j~ 87 ~ @X ~ @X ~ da
dX
: 92
The vector identities in Eqs. (81)(87), can be employed dt @t @a dt
to show that Then the evolution equation for the wake is given by
 
cos a cos b aj jb~jj~
j~ aj jb~j  b~ ~
a ~ @X
@X ~ da
; 88 ~;
U 93
h j~ ~ ~
a bj jbj j~ aj @t @a dt
436 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

cally, arbitrary body shapes may be considered. Panel


methods are usually coupled to a far-wake model or
coupled with a grid-based full potential or Navier
Stokes method (see Section 4.2.7). In most cases, the
disturbance due to the rotor blade is linearized about the
free stream. In what follows, we consider the case of
incompressible ow although the extension to linearized
compressible ow is straightforward [60]. The use of
panel methods in computational uid dynamics is
reviewed by Hess [61] and applications to rotary wing
aerodynamics are discussed by Morino and Gennaretti
[62] and by Caradonna [1].

4.1.9.1. The lifting line. In vortex representations of the


rotor wake, the blade is often represented by a lifting
line of bound vorticity. The lifting-line model is the rst
term in a matched asymptotic expansion in the inverse
aspect ratio 1=A 2c=b where c is the chord of the
rotor blade and b=2 is its length [63]. Rotor blades
typically have aspect ratios of the order of 20; for model
blades used in rotor tests, the aspect ratio is much
Fig. 16. Geometry for a relaxation method. smaller.
Because vorticity is conserved, any variation of bound
circulation must result in shed and trailing circulation
where U ~ is the velocity eld induced by all the (Figs. 1214). For a lifting line, if the bound vortex is
components in the wake; the vortical portion of the assumed to lie in the plane z 0; the imposition of the
velocity eld may be calculated using the BiotSavart vanishing normal ow condition leads to, for a xed
law. Here X ~ x; y; z: In practice, t z=O and a wing [32],
Z !
cb =O where cb and z are dened in Fig. 16 and in this
1 b=2 dG dy0
case da=dt 1 since z and cb measure the same angles. Gy mc UN a  ; 95
4p b=2 dy0 y  y0
Note that cb is the angle the blade makes with the x-axis
and is a space-like variable, while z corresponds to the where G is the circulation, UN is the speed at N relative
age of a given point on the vortex and is a time-like to the wing; also c is the chord. Here m is the lift-curve
variable. Spatial periodicity is assumed in cb : slope which depends on the shape of the wing and for a
~cb ; z X
X ~cb 2p; z at-plate, rectangular wing, m 2p: This is the linear-
ized lifting-line equation of Prandtl.
and a periodic initial condition is also assumed. The Unlike the bound circulation on the xed wing, the
equation thus becomes bound circulation on a rotary wing increases approxi-
~ @X~ U
~ mately linearly with span and the lifting-line equation
@X
: 94 for a rotary wing is obtained by setting UN Oy for
@z @cb O hover (with obvious generalization to forward ight,
Eq. (94) is discretized using standard explicit or implicit ascent, etc.) in Eq. (95) [29]:
methods (Euler, modied Euler) and given an initial Z !
1 b=2 dG dy0
wake distribution, the scheme is iterated to a nal steady Gy mc Oya  : 96
solution. 4p b=2 dy0 y  y0

However, the angle of attack must now be viewed as


4.1.9. Representation of the blade including the eect of the rotor wake which moves down
There are in general three ways to represent a rotor away from the blades, and so
blade: (1) as a lifting line; (2) as a zero-thickness lifting
surface and (3) as a nite-thickness airfoil. Lifting-line a a0 ai a0 1  wi y; 97
and lifting-surface methods for the representation of the where ai is the induced angle of attack and wi is the
blade are special cases of a more general class of induced downwash. Note that for the wing of large
methods referred to as panel methods, which can be aspect ratio we have
used to represent the thickness portion of the airfoil. The
primary advantage of panel methods is that theoreti- GBmcOya0 1  wi y: 98
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 437

blade. From the BiotSavart law, Eq. (57), the vertical


velocity induced by the wake on the surface is given by
wx; y; 0; t
Z
1 gy x  x0 gx y  y0 
 dx0 dy0 :
4p S x  x0 2 y  y0 2 z  z0 2 3=2
99
Here gy is the circulation per unit span aligned along the
blade, gx is the circulation per unit chord aligned in the
x-direction in the wake and S denotes the surface which
comprises the blade and the wake. Thus, the equation
that must be solved for the bound circulation is for a
xed wing [32]:
Z
1 gy x  x0 gx y  y0 
 dx0 dy0
4p S x  x0 2 y  y0 2 z  z0 2 3=2
UN sin a0 0; 100

Fig. 17. Sketch of the bound circulation on a typical rotor where a0 is the geometric angle of attack. The linearized
blade in hover. version of Eq. (100) is obtained by assuming the wing is
innitesimally thin and sin aBa and thus
Z
1 gy x  x0 gx y  y0  0 0
 dx dy UN a0 0
4p S x  x0 2 y  y0 2 3=2
Thus, to leading order, the bound circulation is
approximately linear as shown in Fig. 17. Of course, 101
Eq. (98) is not valid close to the blade tip where the and we have linearized the problem so that the wing is in
circulation must drop to zero. It should also be noted the plane z 0: Similar to the case of the lifting line, for
that the shape of the bound circulation curve depends on the rotary wing we write UN Oy in hover and a
blade shape, rotation speed, ight speed and other a0 1  wi x; y; 0; t and
factors. Z
1 gy x  x0 gx y  y0  0 0
At higher angles of attack, the blade may be near stall  dx dy
and the lift coecient must often be obtained from 4p S x  x0 2 y  y0 2 3=2
experimental data or from a simplied lifting model [64]. Oya0 U ~B . n#; 102
Airfoil lookup tables are also used to provide the lift and
where U~B is the velocity of the blade which accounts for
drag coecients. The static stall angle for many rotor
blades is generally in the range of 12151 although trim and aeroelastic blade motions. If we separate out
dynamic stall angles may be above 201: The strength of the eect of the wake, the strength of which is known
the tip vortex is somewhat a matter of conjecture and its from the Kutta condition, Eq. (102) can be written as
Z
value is usually specied from experiment or taken to be 1 gy x  x0 gx y  y0  0 0
 dx dy
the maximum of the bound circulation on the blade [65]. 4p SB x  x0 2 y  y0 2 3=2
Oya U ~B . n#; 103
4.1.9.2. The lifting surface. Lifting-line theory is simple
and fast computationally and it has often been stated where a a0 1  wi x; y; 0; t is the induced angle of
that lifting-line theory will fail if large spanwise attack and wi is the induced vertical velocity at the rotor
variations in velocity on the blade occur. This assertion disk, i.e. the downwash or inow. Note that on the blade
has been called into question [75,76]; in that work g gy and in the wake, at least for a xed wing g gx :
lifting-line and lifting-surface analyses of ow past xed The integral equations given just above can be
and rotary wings respectively give the same result to discretized and solved for directly. Most often, the
within numerical error. Nevertheless, most rotor design blade is viewed as a collection of straight line
codes use an approximation known as lifting-surface vortex segments with the velocity distribution given in
theory [32,66] in which the blade is replaced by a single Section 4.1.7. In each case, the solution would proceed
at surface of innitesimally small thickness. As with the as follows. The blade is divided into N surface panels
lifting line, the integral equation for the strength density and the wake is discretized into a series of vortices as
of the vortex sheet that represents the wing, and the indicated in Fig. 12. As mentioned above for a two-
vortex sheet that represents the wake is determined by dimensional airfoil, each panel has been replaced by a
imposing the solid wall boundary condition on the two-dimensional point vortex (i.e. the lumped vortex
438 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Fig. 18. Vortex ring model for a rotor blade as a lifting surface. Fig. 19. Horseshoe vortex model for a rotor blade as a lifting
surface.

element), while for a three-dimensional blade, in a low- where wi; j is the dimensionless induced downwash due
order method, each surface panel is represented by a to the presence of the wake at the jth panel. Here Gk is a
three-dimensional vortex line of constant circulation on one-dimensional array of vortex strengths and depends
each panel. There are several vortex congurations on the numbering system; for example, the rst N
which can be chosen. In Fig. 18, each panel is elements of G would generally correspond to the rst
represented by a vortex ring on its boundary. In this row of panels near the leading edge of the rotor blade.
case the collocation point at which the boundary Also m M N is the number of panels and M is the
condition is satised can be taken to be located at the number of panels in the streamwise direction and N is
geometric center of the panel. An alternative is to the number of panels in the spanwise direction. The last
represent each panel by a vortex whose circulation is row of the above set of equations corresponds to the
oriented in the spanwise direction along the blade Kutta condition. Consider the case of a horseshoe
at the 1=4 chord line at the spanwise center of the panel. system on the blade, then gx 0 in Eq. (101) and
In this case the wake may be represented by a series of Gk gyk Dyk :
horseshoe vortices ([67], Fig. 19) and the boundary In particular in hover, the wake circulation is constant
condition is satised at the 3=4 chord line. Recall that along the wake vortex lament and in steady ow
for the two-dimensional symmetric airfoil, the center of computations, assuming periodicity, the bound circula-
pressure is at the 1=4 chord. Note that placement of the tion is solved for and then the wake position is iterated.
collocation point at the 3=4 chord line gives the proper The wake elements may be chosen to consist of vortex
lift for a single vortex placed at the 1=4 chord line as boxes or horseshoe vortices. The linear system of
demonstrated for two dimensions ([33], p. 262). equations for the bound circulation may be solved by
Thus the panel circulations Gk ; k 1; y; m may be standard numerical techniques such as Gauss elimina-
evaluated from the surface boundary condition, ex- tion, direct matrix inversion, or for a large set of
pressed as the set of linear equations equations, say GaussSeidel.
X
m
Ajk Gk 4.1.10. Potential theory methods
k1
Generalized panel methods based on potential theory
~Bj . n#;
a0 1  wi; j U j 1; y; m; 104 have been used in rotor codes for a number of years.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 439

These generalized methods originate from the three- rfx; y; z; t


dimensional representation theorem for any solution Z " !#
of Laplaces equation which for incompressible ow 1 1 @ 1
 sr mr dS
is [68] 4p SB ~ X
jX ~0 j @n jX
~X ~0 j
Z !
fx; y; z; t @ 1
m r dS rfN 108
Z " !#
SW @n ~X
jX ~0 j
1 1 @f @ 1
f dS fN ;
4p S jX ~0 j @n
~ X @n jX
~X ~0 j
105 and dotting this equation with the normal to the blade
we obtain
where n is the coordinate normal to the rotor blade and
Z " !
f is the velocity potential. Dening
~B  1
n# . U sn# . r
1
@f 4p SB ~ X
jX ~0 j
s !#
@n
@ 1
mn# . r dS
and m f we have @n jX
~X ~0 j
Z " Z !
1 1 @ 1
fx; y; z; t  s mn# . dS n# . rfN : 109
4p ~ X
jX
S ~j 0
SW @n jX
~ X~0 j
!#
@ 1
m dS fN : 106 This is the Neumann problem in which the velocity
@n jX
~ X~0 j
gradient normal to the surface is specied. In this case,
Here m is termed the doublet distribution and s the the source distribution is known and the doublet
source distribution since the velocity due to a source in distribution on the surface is unknown.
three dimensions is To solve the integral equation (109), we represent the
! blade by a number of sources and doublets and obtain a
~ 1 set of linear equations of similar form as those given
V source sr
~X
jX ~0 j above for the vortex lattice. In the present case

and for a doublet X


N X
N
! UBj UBk Akj mBk Bkj wj wN; j
~doublet @ 1 k1;kaj k1;kaj
V mr :
@n jX
~ X~0 j j 1; y; N 110

The surface S again consists of the blade and the


wake: typically, the blade is represented by both which is a system of N equations in the unknown
source and doublets while the wake is represented doublet strengths mBk given the known downwash wj :
by doublets since the doublet distribution is equivalent Here Akj and Bkj are coecients representing the
to a vortex representation of the wake [32]. Thus, we integrand of each of the integrals in Eq. (109). For
write lifting surfaces such as the rotor blades themselves,
doublets are often employed exclusively, while for non-
Z " !#
1 1 @ 1 lifting surfaces, such as airframes, sources are often used
fx; y; z; t  s m dS exclusively. As with the vortex method discussed just
4p SB ~X
jX ~0 j @n jX~X ~0 j
Z ! above, the velocity boundary condition is satised at a
1 @ 1 given control point on a panel. The wake doublet values
 m dS fN : 107
4p SW @n jX ~ X ~0 j in the expression for the downwash (inow) wj are
known by application of the Kutta condition and are
The velocity potential anywhere in the oweld is given by the dierence in the potentials fU  fL at each
determined by the values of the potential and its spanwise location at the last chordwise panel; for steady
derivatives on the surface of the blade. The above ows the wake doublet values are constant along
solution is written for a xed blade but the analysis can streamlines. In the unsteady case, the wake doublet
be extended to a blade undergoing arbitrary motion values vary both streamwise and spanwise.
[69,70]. It is also possible to specify the potential on the
To apply the boundary condition on the blade we surface of the rotor blade. Consider the direct discretiza-
dierentiate Eq. (107): tion of Eq. (107). Then if we specify the velocity
440 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

potential at a given point j to be zero where V~ is the linear speed of the aircraft. The pressure
coecient can be obtained as
X
N X
N
4pmj UBk Akj mj Bkj  2pmj pN  p
cP ;
k1 k1;kaj rOR2
X
Nw
mwk Ckj fNj ; j 1; y; N: 111 where OR is the tip speed. To obtain the lift and induced
k1 drag, the pressure can be evaluated at the collocation
points and integrated over the surface and in dimension-
Eq. (111) is the discretized version of the external less form:
Dirichlet problem and variations of this method have Z
been used in most rotor codes [71]. In some methods, the Fn cP dS: 114
external Dirichlet problem has been combined with the S
internal Dirichlet problem to eliminate the left hand
Any one of a number of integration schemes may be
side of Eq. (111) as well as the term involving fN :
used to evaluate the integral.
These so-called panel methods require that the
The lift and drag can also be calculated using the
distribution of the strengths of the source or doublet
KuttaJoukowsky law. Consider the case of a vortex
be specied. If these strengths are assumed to be
lattice model of the blade and wake. The force induced
constant on each panel, the method is referred to as
by a three-dimensional patch of vorticity is given by
low order. High-order methods use a linear or quadratic
([74], pp. 4648).
distribution of the strength over the panel [71,72]. A list Z
of panel methods and the types of singularities they use F~ rU~ o~ dV;
is given by Katz and Plotkin ([32], p. 402). V

where o ~ is the vorticity and the integral in this equation


4.1.11. Performance is over the volumetric region where oa0: Consider rst
Recall that the primary function of a wake calculation the blade as a lifting line. If we linearize about the local
scheme is to provide input to calculate the total thrust free stream velocity UN Oy; and if we assume that the
generated by the rotor. Typically, the thrust required for bound vortex system on the blade has a vorticity
operation is given and the length and chord, as well as component in the y-direction, then each panel has a
the overall geometry of the blade (i.e. twist, camber) are section lift corresponding to
designed to give the required thrust. This design
procedure requires that the inuence of the wake on dF~ rGb Oyi# dy j#;
the overall momentum through the rotor disk must be
where Gb is the strength of the bound vortex, and i# and j#
determined. This means that given the thrust, the
are unit vectors in the coordinate directions x and y:
geometry of the wake must be determined; from the
This is precisely the KuttaJoukowski law. On the other
wake geometry, the lift and drag coecients and
hand, in the general case
the downwash can be calculated and the actual thrust
(weight the downward force due to the wake) required dF~ rU ~V Gx dx i# Gy dy j# Gz dz k#:
~N U
is known. Recall that the thrust and power coecients 115
depend on the lift and drag coecients.
While the calculation of performance parameters may Here U~V is the velocity distribution induced by both the
be done in a variety of ways, it is useful to describe bound vortex distribution on the blade and the vortex
briey the nature of such calculations. First, the pressure system that makes up the wake. Gx is the circulation of
is calculated from Bernoullis equation, which in an a vortex whose axis is aligned with the x-axis. In linear-
inertial reference frame is ized analyses, Gz 0 on the blade. Eq. (115) can be
"      # integrated to obtain the force
pN  p 1 @f 2 @f 2 @f 2 @f
: 112 X
N
r 2 @x @y @z @t F~ r ~V;k Gxk Dxk i# Gyk Dy j#: 116
~N;k U
U
k1
In a coordinate system rotating with the blade, the
relative velocity must be used and the Bernoulli equation In particular, for the horseshoe vortex model, Gx 0:
will have the form [73] In this case, the lift,
"      # Z X N
pN  p 1 @f 2 @f 2 @f 2 L dFz r Gyk UN;k UVxk Dyk ; 117

r 2 @x @y @z k1

~ ~
~O @f where Dyk is the grid spacing and as before N is the
 V r . rf ; 113
@t number of points on the blade. Similarly, the induced
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 441

drag for this case is given by


X
N
Di r Gyk wk Dyk ; 118
k1

where wk is the induced velocity due to all the wake


components evaluated on the lifting line. In particular,
in the present case of an innitesimally thin blade, wk
UVz;k : Note that the lift is generated predominantly by
the bound and shed vortex segments while the induced
drag is induced predominantly by the trailing vortex
system.
In the same way, we can obtain the aerodynamic
moment for the lifting line as
~ ~
dM y Gb Oyi# U
y dF~ r~ ~V dy j#: 119
Integrating, we have
X
N
~
M yk F~k ;
~ 120
k1

where F~k is the force vector at each segment of the


blade.
For the lifting surface, we have for the case of a
horseshoe vortex model for the blade,
Xn
Lr gyk UN;k UVx;k DAk ; 121 Fig. 20. Classical or rigid rotor wake and an actual wake based
k1 on experimental results for a single blade in hover as described
by Landgrebe [52].
where n is the number of panels, DAk is the area of each
panel and gy;k is the circulation per length of each panel.
The induced drag is given by features of the inow are reproduced, while amplitudes
X n of the doublet spike are underpredicted for the fully
Di r gyk wk DAk : 122 trimmed case and for the untrimmed case near 75%
k1 span. The authors note that this phenomenon is due to
The lift and induced drag coecients are dened by the fact that only 24 harmonics are employed in the
L eigenfunction expansion.
cL 1 2 Note that these potential ow methods do not give the
2 rOR A viscous portion of the drag and this quantity must be
and obtained by other means.
Di
cDi 1 2 4.1.12. Results for the rotor wake
2 rOR A 4.1.12.1. Rigid and prescribed wakes. Having discussed
and compressibility is incorporated via the Prandtl the basic structure of the rotor wake and the representa-
Glauert relation tion of the rotor blades, the precise position of each of
cL Ma 0 the tip vortices shed from each blade and the position of
cL Ma > 0 p ; each vortex sheet can be calculated. In modeling the
1  Ma2 rotor wake by systems of vortices, three approaches
where Ma is the Mach number. have generally been used: rigid wake models prescribed
A novel approach for calculating the inow at the wake models and free wake models. In the rigid wake
rotor disk has also been developed [7779] based on an model, the vortex system position is specied as a
eigenfunction approach. While a lift model is required function of advance ratio and thrust; this situation is
the method is easy to couple with an aeroelastic model compared with results from experiment in Fig. 20 [52].
for the rotor blade. For the unsteady case where trim is The rigid wake model does not account for contraction
incorporated, the calculations are reduced to ordinary of the wake as is predicted by momentum theory and
dierential equations in time for the coecients in the thus blade load calculations often disagree signicantly
eigenfunction expansion for the pressure. Comparisons with experiment as rotor solidity, thrust level and tip
with experimental data [79] show that the essential Mach number are increased [52].
442 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

The prescribed wake model [52,53] employs experi-


mental data to locate the vortex positions and thus
incorporates wake contraction. Because experimental
data are used to locate the wake, this wake method is
very ecient computationally; however, the method
cannot be extended to parameter ranges and ow
conditions for which there is no experimental data.
The prescribed wake concept has been extended to
numerical prediction of rotor airloads for a range of
ight conditions [80].

4.1.12.2. Free wake computational schemes. In a free


wake calculation, the vortex system motion is calculated
directly from the eects of all the other wake compo-
nents and the inuence of the blade. In this method the
wake is allowed to develop in time and initial value Fig. 21. The calculation of a spatially and temporally periodic
Lagrangian methods or relaxation techniques are used rotor wake by Miller and Bliss [57]. Only the tip vortex is
to determine its position at each time step. The free wake shown.
is the industry standard today and as computers have
become faster, the free wake calculation has become
more aordable. presented in Fig. 21. They were also able to obtain
There are three aspects of free wake methods which results for forward ight. A relaxation technique which
seem to distinguish dierent schemes: rst, whether a uses curved vortex elements to evaluate hover perfor-
lifting line or lifting surface model is used for the blades; mance has also been described [54,58]. The blade is
second, whether relaxation methods or a time advance represented as a lifting surface. Compressibility eects
method is used; and third, the form of the discretization are incorporated and the drag coecient is obtained
of the BiotSavart law used. Various complex phenom- from two-dimensional airfoil data and results for a
ena observed in propeller and rotor wakes have recently number of rotors are presented. A typical result is
been discussed [81]. depicted in Fig. 22 [58]. Calculation of the wake of a
The vast majority of the calculations which have four-bladed rotor in forward ight using a pseudo-
appeared in the literature are for hover and forward implicit iterative technique has also been described [59].
ight. Current wake calculations based on a vortex A comparison of several methods for computing
method can be classied as research modules or design wake-induced airloads [82] and the results are depicted
modules based on whether the wake calculation is in Fig. 23 at an advance ratio 0:1: Figs. 23(a)(c) depict
coupled with a comprehensive rotor design code. Most results for a vortex lattice approach using straight line
of the research calculations focus on developing new segments to discretize the entire wake (Fig. 23(a)), a
methods for predicting the wake geometry; performance rigid inner wake and a free tip vortex (Fig. 23(b)), and a
calculations may or may not be made. Wake calcula- curved vortex line approach based on the constant
tions which are coupled to a rotor design code must, by vorticity contour approach, respectively [45]. Note the
necessity, use simpler and computationally faster meth- considerable dierences in the wake geometry for each
ods to predict the rotor wake. of the three methods. The dierences are particularly
The rst free wake hover calculations appear to be acute near the blade tips where the vorticity eld is
time-marching calculations [47] followed closely by a strongest. For Figs. 23(a) and (b) only the tip vortex is
relaxation technique [55,56]; Scully [47] computed shown although the entire wake is calculated; for
solutions for the rotor wake using a time-marching Fig. 23(c) the constant vorticity contours along the
vortex method. Using the swirl velocity given by entire blade are shown. Airload results (not shown)
Eq. (60), he used the straight line segments of Fig. 12 indicate that the constant vorticity contour method
to model the tip vortex. The inboard vortex sheet was compares better with experimental data than the others
represented by a single large-core vortex located at especially at low advance ratio, but all the results for the
midspan. It was found that the tip vortex is the blade loads exhibit signicant deviation from experi-
dominant component of the wake velocity eld because mental data.
its strength is much greater than that of the inboard The presence of physical instabilities in the wake
sheet. resulting in signicant vortex interactions in hover has
Modern relaxation methods are illustrated by the been observed [52,53]. The physical instabilities are a
work in Ref. [57] which uses a linearized version of the large perturbation from the perfect helical form of the
BiotSavart law and the result for the hover case is motion of the tip vortex depicted in Figs. 21 and 22. An
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 443

Fig. 22. Wake geometry for a rotor in forward ight at an


advanced ratio of m 0:08 [45]. Only the tip vortex is shown.

Fig. 24. Wake geometry for a rotor in hover after the rotor has
turned 4481: The vortices do not end but are connected to a
model of the far wake which is not shown [84].

neglected. A typical result is shown in Fig. 24. Note how


the rst three turns of the wake are very regular
although the distance between each individual turn of
the tip vortices is not uniform. The results of [84] are
spatially and temporally periodic, but not axisymmetric.
A relatively small time step corresponding to a rotor
phase angle of 41 must be used to preserve accuracy.
A comparison of the wake position with experiment is
depicted in Fig. 25 and note that the agreement with the
experiment is good. It has also been shown that in
ascent, the irregularities do not appear.
Fig. 23. Comparison of various wake models which utilize The methods described so far involve the calculation
vortex methods to describe the rotor wake [82] at advance ratio of the wake-induced airloads in which the wake vortex
0:1: The tip vortex only is shown. (a) Vortex lattice, rigid blade; system is described in terms of the BiotSavart integral
(b) Scully wake, straight line segments; (c) curved vortex and/or superposition of related elementary singularities.
segments. The inboard vortices are omitted for clarity. As such no computational grid in the wake is required
and for this reason, the methods described here are
normally not viewed as computationally intensive. With
example of the type of irregular behavior had already the advances in speed and memory of computational
been identied experimentally [52,53] and calculations hardware, increasingly larger problems may be solved
showing this behavior have been presented [83]. These with a nite-dierence or nite volume approach in
calculations and more recent work [84] were motivated which the velocities in the wake are calculated directly.
by the experimental results presented by Caradonna These methods are described next.
et al. [85] who show photographs of two tip vortices
shed from dierent blades rolling around each other (see
Fig. 31). Recent, more advanced time-accurate techni- 4.2. Grid-based wake calculation methods
ques have been employed [84] to show that the
irregularities are not a result of numerical error but 4.2.1. Introduction
are a physical phenomenon as suggested by previous The term computational uid dynamics (CFD)
experiments [52,53]. In [84] an Adams Moulton method generally refers to grid-based solutions of the Navier
was used to advance each of the tip vortices; the Stokes equations or some subset of these equations such
inuence of the inboard sheet on the motion of the tip as the Euler equations for inviscid ow or the full
vortex has been shown to be minor and has been potential equation for potential ow. There are a wide
444 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

4.2.2. Grid generation


As a rst step in the calculation of the rotor wake
using a grid-based method, a grid or grids must be set
up. This involves mapping a rectangular domain in
Cartesian coordinates into generalized body-conforming
coordinates; the mapping takes the form
x xx; y; z; t;

Z Zx; y; z; t;

z zx; y; z; t;

t t;
where x; y; z is the coordinate system in physical space.
Fig. 25. Comparison of the experimental and computational Under this transformation, Eqs. (40) become
trajectories of the tip vortex of blade-1 for the experimental
data of [85]. Experimental data are represented by  and # @F# @G
@U # @H #
0; 123
computational data by the solid line. The rotor is two-bladed @t @x @Z @z
and in mild axial ight. One and one-half revolutions of the
rotor are shown. where the # quantities are related to the original vectors
by
P
variety of methods used in the calculation of rotor wakes P# ;
and not every method can be discussed in this review. J
Rotorcraft computations, in general, use methods where J is the Jacobian of the transformation
developed for use in ows over xed wings and so the  
 xx xy xz 
objective of this section is to provide the reader with an @x; Z; z  

informal introduction to the methodologies employed to J  Zx Zy Zz :
@x; y; z  
solve the NavierStokes equations with specic applica- z zy zz 
x
tion to the computation of rotor ows. Additional
There are several dierent ways of generating grids.
details of the methods described here may be found in
Perhaps the simplest is to develop an algebraic
the book by Tannehill et al. [86].
transformation to achieve the desired grid; the grid
It is generally acknowledged that there are ve main
should be ne near the rotor blade and increase in size as
issues involved in the CFD solution of helicopter rotor
the outer boundary of the computational domain is
ows; these are
approached. Grids are also generated by solving partial
* grid generation; dierential equations; elliptic grid generation requires
* method of choice: nite volume or nite dierence; the solution of the three-dimensional Laplace equation
* discretization of the convective terms; for the given geometry. The idea is that since Laplaces
* algorithm used to solve the discretized equations; and equation preserves orthogonality, lines of constant-
* specication of initial and boundary conditions. dependent variable (x; Z; z in this context) will be
orthogonal, thus giving a suitable grid in the trans-
Many of the methods used in CFD calculations of the
formed plane. Extra control over the grid is obtained by
rotor wake have been taken from the xed-wing
the inclusion of a right hand side, in which case a
community. Moreover, most NavierStokes codes have
Poisson equation is solved [86].
the ability to turn o the viscous uxes in order to solve
On the other hand, hyperbolic grid generation is
the Euler equations. In keeping with tradition, vectors in
appropriate when the external ow about a single body
this section are denoted by bold type.
is considered as in the applications of interest here [86].
The solution of the compressible NavierStokes
The solution is marched outward in space toward the
equations is particularly dicult in a rotorcraft context.
outer boundary of the computational domain. The three
Near the blade tip, the ow can be supersonic and so
sets of characteristics in a three-dimensional problem
shock waves may be present. Special methods are
yield the appropriate grid.
required to resolve the position and strength of the
In helicopter aerodynamics applications, at least ve
shock waves; the presence of shock waves requires that
types of grid systems have been used depending on the
the convective terms in the equations should be
domain of interest:
upwinded. On the other hand, away from the immediate
vicinity of the blade tip, the ow may be substantially * C-grid;
incompressible. * O-grid;
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 445

* H-grid; time derivative, a rst-order explicit formula is given by


* overset grids; and
* unstructured grids. # n1  U
U # n Dt RHSn ; 124
ijk ijk

The C- and O-grids are used for computations in where RHS refers to all of the spatial derivatives in the
which the primary objective is to calculate the ow on governing equations. The error in this approach is
and around the blade. H-grids are employed if the entire ODt2 over a single time step. The explicit approach
wake of the blade or blades is of interest. Overset approximates the equation at xi ; Zj ; zk at the time
(Chimera) grid systems use a set of several dierent grids interval n and using central dierencing for the
to solve for the ow variables in dierent regions. In derivatives, Eq. (123) becomes
general, there are at least three sets of grids: a body-
conforming grid to solve for the ow near the blade; an # n1  U
# n aF# n #n
U ijk ijk i1=2; jk  Fi1=2; jk
intermediate grid to provide transition to the far eld:
#
bG
n
# n
# n
# n
and a standard Cartesian grid which is used to compute i; j1=2;k  Gi; j1=2;k gHi; j;k1=2  Hi; j;k1=2
the ow velocities far from the blade. Interpolation n n
#
aF# vi1=2; jk  F# vi1=2; jk bG
n
# n
vi; j1=2;k  Gvi; j1=2;k
methods are used to transfer information between grids.
#
gH
n
# n
Overset or Chimera grids can simplify the ow in the vi; j;k1=2  Hvi; j;k1=2 ; 125
neighborhood of the blade which may be undergoing
complex motions in order to trim the rotor in forward where
ight [87]. Overset methods are generally viewed as
being more accurate than conventional methods at the Dt Dt Dt
a ; b ; g :
cost of slightly higher CPU time [88]. Combinations of Dx DZ Dz
the above grids have also been used.
Another type of grid system that is often used is the To preserve second-order accuracy in time a three-point
unstructured grid. The unstructured grid allows ecient backward dierence scheme has also been used:
grid generation around complex congurations [89].
n1
Moreover, the unstructured nature of the method allows @u 3ui; j  4uni; j un1
i; j
: 126
easy insertion and deletion of mesh points, thus @t 2Dt
permitting grid adaptation in a computationally ecient
manner. Unstructured mesh cells usually take the form Even with a three-point backward dierence approx-
of tetrahedral elements and the mesh generation codes imation, the accuracy of the computation is limited by
usually have the ability to subdivide the elements if the fact that the global error after a number of time steps
necessary. is only ODt: Moreover, there are also stability
An example of a C-grid is depicted in Fig. 26(a) [90]. constraints that require that the time step be unreason-
Note that the C-grid is body conforming and is useful ably small. This fact increases the computation time
for calculating the ow velocities near the blade. signicantly and so, in practice, explicit methods are
Combinations of these grid systems are usually em- used only as predictor steps in rotor wake computations
ployed as well. Fig. 26(b) is an example of a CH-grid if at all.
used by Srinivasan et al. [91]. The H-grid allows The alternative is to use an implicit discretization
computation of the ow variables deep into the wake. which leads to a system of linear equations which need
An unstructured grid is depicted in Fig. 26(c) [89]. to be solved at each time step. The idea of the implicit
It should be pointed out that many rotor codes scheme is to approximate the time derivative at the
employ grids that are in motion. In this case, the uid midpoint time level n 1=2: Using a central dierence
velocities in the governing equations are those velocities approximation for the time derivative, we write symbo-
relative to the moving grid. lically in physical space
Once the grid is constructed, the discretization scheme
must be selected. We now consider the basics of the U # n DtRHSn RHSn1 :
# n1  U 127
ijk ijk
central dierence approximation to the governing 2
equations.
The error in this approach is ODt3 over a single time
step. To illustrate the implicit approach, consider the x-
4.2.3. Finite-dierence methods momentum equation of the two-dimensional form
Finite-dierence methods approximate the derivatives of the NavierStokes equations in the thin-layer
in the governing equations by formulas obtained from approximation. Then using central dierence approx-
the truncation of a Taylor series valid locally around a imations for all the derivatives except for those in the
mesh point. Using a forward dierence formula for the x-direction of the convective terms we obtain in the
446 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Fig. 26. Examples of grid systems used in CFD analyses of the rotor wake: (a) C-grid [90]; (b) CH-grid [91]; and (c) unstructured
grid [89].
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 447

physical space, In vector form we thus have


  Z Z
m Dt @U
n1 Dt dx dy dz r . M dx dy dz 0; 130
ui;n1 v
V @t
j1 i; j
2Dy 2Dx2 V
 
n1 n1 Dt m Dt where
ui; j 1 ui; j
2Dx Dx2
  M F; G; H:
n1 Dt m Dt
un1
i; j1 vi; j RHS; 128
2Dy 2Dx2 We now apply Gausss theorem to obtain
% Z
dU
where we have assumed that the convective term M  n dx dy 0; 131
! dt A
n1 n1
@u ui1; j  ui; j
u un1
i; j ; where n is the outward unit normal to the surfaces of the
@x Dx cell volume and
Z
that is, we use a rst-order forward dierence and have U% 1 U dx dy dz
assumed that un1i; j > 0: More will be said about upwind
V V
dierencing in the next section. For xed i; that is, at a
is the average value of U over the cell volume.
given streamwise location, Eq. (128) is a set of equations
To illustrate the discretization scheme, consider the
for the unknowns ui;n1 n1
j1 ; ui; j and un1i; j1 : They are single two-dimensional scalar equation
nonlinear since both ui; j and vn1
n1
i; j appear in the
coecients. To linearize these equations we assume that @U @F @G
0:
both un1 n1
i; j and vi; j are known in the expressions for
@t @x @y
the coecients and then iterate. Thus, initially within the
Then the corresponding nite-volume equation is given
iterative cycle, we set un1 n n1 n
i; j ui; j and vi; j vi; j in the by
coecients of the linearized system. Similar equations Z Z
appear in the other two coordinate directions. The d
U dx dy M  n ds 0; 132
linearized set of equations is a tri-diagonal system which dt A C
may be solved directly using the Thomas algorithm [86]
where M F; G and n is the outward unit normal to
which in multiple dimensions is called the alternating
the contour. Let us consider a single cell as depicted in
direction implicit (ADI) method. A popular variation of
Fig. 27. The center of the cell is labeled i; j: Then at the
the ADI method for compressible ows is that of Beam
center of the cell, if we approximate the line integrals
and Warming [92]. If we isolate un1 i; j on the left hand around the boundary of the cell by the trapezoidal rule,
side of Eq. (128), then the set of equations could be
we obtain
solved by an iterative method such as GaussSeidel or
successive over relaxation (SOR). Both types of methods dU% i; j 1
Dx Dy DyFi1=2; j1=2 Fi1=2; j1=2
are used today as well as a LU-decomposition in dt 2
combination with the GaussSeidel method [93,94] and 1
DxGi1=2; j1=2 Gi1=2; j1=2
other similar methods. 2
In general, a computational scheme will proceed as 1
DyFi1=2; j1=2 Fi1=2; j1=2
follows. The velocities and density are calculated from 2
the momentum equations and the continuity equations. 1
DxGi1=2; j1=2 Gi1=2; j1=2 : 133
The total energy which is made of internal energy and 2
kinetic energy is calculated from the energy equation.
The error in this equation is the maximum of ODx2
The total energy minus the kinetic energy yields the
and ODy2 : If the values of F and G are known on the
internal energy, which then gives the temperature.
boundary of the cell, then Eq. (133) gives the time
Finally, the pressure is calculated from the equation of
evolution of the velocity at the center of the panel. If left
state: p rRT:
in this form, the method is called a vertex-based nite-
volume method. Other more accurate integration
4.2.4. Finite-volume methods schemes are often used, including Simpsons rule and
The nite-volume method uses the integral form of NewtonCotes sloping integration formulas.
the governing equations. For simplicity, we work with We can also write the line integrals in terms of the cell
the untransformed grid system and integrate Eq. (40) centered point as well. Now, from Taylor series, for a
over a volume V: constant grid size, the standard second-order interpola-
Z   tion formula yields
@U @F @G @H
dx dy dz 0: 129 1
V @t @x @y @z 2Fi1=2; j1=2 Fi1=2; j1=2 Fi1=2; j 134
448 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Fig. 27. Sketch of a rectangular cell in two dimensions. The


two vertices to the left of the cell may be used in a backward
interpolation formula in an upwind, nite-volume scheme.

and
1
2Gi1=2; j1=2 Gi1=2; j1=2 Gi; j1=2 : 135
Furthermore, by the same reasoning, Fig. 28. Sketch of a rectangular cell in three dimensions.
Gi; j1=2 12Gi; j1 Gi; j 136
and so Eq. (133) becomes
on the one face, and
dui; j 1 Z
Dx Dy DyFi1; j  Fi1; j
dt 2 F dA Fi1=2; j;k Dy Dz
1 A
DxGi; j1  Gi; j1 0:
2 on the other face. In a cell centered scheme, we can
Dividing by Dx Dy we have interpolate to obtain Fi1=2; j;k and the simplest formula
dui; j 1 1 is the second-order accurate simple average given by (see
Fi1; j  Fi1; j Gi; j1  Gi; j1 0 Fig. 28)
dt 2Dx 2Dy
137 Fi1=2; j;k 12Fi1; j;k Fi; j;k :
which we recognize as the central dierence approxima- In this case, in scalar form a single governing equation
tion to the governing equations. However, we note becomes
that values of F and G are the averages over the cell.
dui; j;k 1
Formally, then, nite-dierence and nite-volume meth- Dx Dy Dz Dy DzFi1; j  Fi1; j
dt 2
ods are equivalent.
1
In three dimensions, the integral form of the govern- Dx DzGi; j1  Gi; j1
2
ing equations takes the form
Z Z Dx DyHi; j1  Hi; j1 0:
d
U dx dy M  n dA 0; 138 Dividing through by the volume, we obtain
dt V A

where now M F; G; H and n is the outward unit dui; j;k 1


Fi1; j;k  Fi1; j;k
normal to a given face. The three-dimensional picture is dt 2Dx
depicted in Fig. 28. The area integral in Eq. (138) can be 1
Gi; j1;k  Gi; j1;k
interpreted as the ux of a given quantity through a 2Dy
surface. In a cell centered scheme, each area integral is 1
the product of the function value at the center of the cell Hi; j1;k  Hi; j1;k 0 139
2Dz
face, times the area, just as in the two-dimensional case.
which is again formally equivalent to the central
Thus, assuming that the cell center is labelled i; j; k then
dierence approximation to the governing equations.
for a scalar ux F;
Z In a vertex centered scheme the area integrals would be
F dA Fi1=2; j;k Dy Dz approximated by, say, the average of the functional
A values at the corners.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 449

In practice, more accurate schemes are required; schemes. However, it is important to note that the
moreover, to incorporate the directional nature of the schemes discussed below which have the primary
convective terms and to properly resolve shock waves purpose of resolving shocks will not work in regions
(the Euler equations are hyperbolic) directional where the ow is incompressible.
dierencing is usually used. For example, a backward Before implementing a ux-splitting scheme, the
interpolation formula may use vertices behind the cell in system of equations is diagonalized. To do this, consider
Fig. 27. This topic is discussed next. the single scalar ux F: We note that F AU and A
is the Jacobian matrix dened just below. Before we split
the uxes, we decouple the equations by diagonalizing
4.2.5. Vector splitting, shock waves and upwind the matrix A: Consider the matrix T; where T is the
dierencing matrix with columns being the eigenvectors of the
It is well known that much of the discretization error matrix A [99] which form a basis. We thus can write in
that builds up in a calculation is due to the error the case of a single direction, say x;
incurred in discretizing the convective terms in the
A TLT 1 ;
governing equations. The second-order central approx-
imation used above in both the nite-dierence and the where L is a diagonal matrix consisting of the
nite-volume approaches is unconditionally unstable in eigenvalues of the matrix A: We can then split the
the von Neumann sense and so some other means of matrix L into positive and negative parts corresponding
approximating the convective terms in the nite- to the positive and negative eigenvalues and write
dierence approach and some other interpolation F TL L T 1 U A A U F F  :
formula in the nite-volume approach must be used.
The solution is to use upwinding approximations to Now,
reect the fact that the governing equations possess @F @F @U
directionality.
@x @U @x
The compressible NavierStokes equations permit
so that A is the Jacobian matrix A @F=@U: The Roe
regions of ow across which the pressure and velocities
scheme [98] averages the velocity from the left and the
are nearly discontinuous; these regions are termed
right as
shock waves. Finite-dierence methods based on Taylor
series approximations are inaccurate in these regions Uy UL yUR  UL :
and the convective terms in the governing equations
Now
written in conservative form must be treated with care. Z R Z 1
Finite-volume schemes are particularly useful for dF
dF FUR  FUL dy
handling shocks since no derivatives appear in the L 0 dy
equations. so that
In general, shock resolving methods (i.e. shock Z R Z 1
capturing) are usually developed for one-dimensional dU
dF A dy
model problems such as the inviscid Burgers equation. L 0 dy
Z 1
They are then extended to two- and three-dimensional
systems by applying the same method to each individual UR  UL A dy UR  UL A* ;
0
equation. However, since the NavierStokes and Euler
equations are coupled, there is no guarantee that the where A* is called Roes average matrix. By denition
methods that work for a single, nonlinear conservation FL Fj1=2 ; FR Fj1=2 :
law will work for systems of conservation laws and no
In a cell centered scheme, these quantities must
unied theory of the application of these methods to
be related to the cell centered values. In Roes
systems of equations exists [95].
scheme,
The treatment of the uxes in conservative form
depends on the directional nature of the uxes; that is, Fj1=2 Fj A DU
waves from both upstream and downstream may pass
and
through an individual cell volume. It is for this reason
that central dierencing, which includes no directional Fj1=2 Fj1  A DU:
preference cannot be used near discontinuities. Thus
Adding these two equations,
the uxes are split into upstream and downstream
running components [9698]. Common splitting meth- Fj1=2 12Fj Fj1  jAjDU; 140
ods include StegerWarming splitting and the van Leer where
splitting in conjunction with nite-dierence schemes,
while the Roe scheme is used with nite-volume jAj A jA j:
450 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

For the simple case of one-dimensional ow of a perfect


gas, the diagonalized matrix is
2 3
U 0 0
6 7
L4 0 Ua 0 5:
0 0 Ua
Once the A matrix is diagonalized, the set of equations
is de-coupled and each equation can be solved sepa-
rately. In the nite-volume formulation and to rst
order in space, uj is the constant average value in the
cell. In eect, then the Roe scheme solves the Rieman
problem [98] which is the classical inviscid Burgers
equation subject to discontinuous initial conditions. For
the one-dimensional Burgers equation A U and
jAj 12UL UR :
There are many dierent methods used to calculate the Fig. 29. Qualitative sketch of the pressure distribution across a
shock showing oscillations which may occur in the absence of
left- and right-hand state variables UL and UR and this is
specialized techniques to eliminate them.
what distinguishes many methods. For example, many
authors have used a scheme developed by Van Leer [100]
known as monotone upwind-centered schemes for
conservation laws (MUSCL):
  i 1=2 may be computed using the points, i; y; i  5; or
d
UL; j1=2 1 1  dkr 1 dkD Uj ; points i 1; y; i  3: The coecients of the polynomial
4
are obtained by matching the uxes at all of the
  interpolation points. Away from shocks, a centered
d
UR; j1=2 1  1  dkr 1 dkD Uj1 ; formula may be used. This method has been shown to be
4
very accurate, more accurate than the third order
where d is a ux limiter and k indicates the order of the MUSCL approach [103].
scheme; k 13 gives third-order accuracy. Here r and D ENO schemes are more computer time consuming
are the backward and forward dierence operators considering the extra calculations performed. Moreover,
respectively: oscillations on the order of the truncation error can still
rUj Uj  Uj1 occur [95]. Local variations in the pressure should look
like the smooth curve in Fig. 29. Another condition
and which can be used in conjunction with the Roe and the
DUj Uj1  Uj : ENO scheme is the total variation diminishing (TVD)
procedure [104]. A procedure is said to be TVD if
All discretization schemes which are employed for the
convective terms in the compressible NavierStokes TVF n1 pTVF n ; 141
equations are susceptible to oscillations near shocks; this where
situation is illustrated in Fig. 29. Various modern
X
jN
techniques have been used to avoid unwanted oscilla- TVF n n
jFj1  Fjn j:
tions. One such method is the essentially non-oscillatory jN
scheme (ENO) [101,102] which has been applied
Sucient conditions for a scheme to be TVD are given
primarily to nite-volume schemes. As applied to the
by Harten [104].
NavierStokes equations, the ENO method chooses
The aforementioned decomposition and smoothing
from several interpolation stencils around the mesh
schemes are used with nite-volume methods. Separate
point at which the ux is required. The interpolating
dierencing methods are used with nite-dierence
polynomials can be either forward, backward or
schemes. Recall that in primitive variable form, rst-
centered [103]. The interpolating polynomials dier in
order upwind dierence scheme in the x-direction, for
what set of points are being used. For example, fth-
example, is dened as
order accuracy may be obtained with ve points in 8
addition to the point at which the ux is required. ENO >
>1 uj1  uj
schemes choose a given stencil based on a condition @u < 2uj jujj Dx
for uj > 0;
u 142
which minimizes the sum of the absolute values of the @x > > u  uj1
: 12uj  jujj j for uj o0:
derivatives around the point. For example, FL at a point Dx
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 451

In todays rotorcraft codes, rst-order dierencing is not


accurate enough and researchers have investigated the
use of higher order stencils [105,106]. In particular,
Wake et al. [106] have shown that ninth-order schemes
are required to convect a vortex accurately over the time
range required for use in rotor design. The increased
order stencils are derived from Taylor series and the
backward formulas are depicted in Table 1. The forward
stencils are obtained by multiplying each coecient by
1: In particular, the third-order upwind scheme in the
x-direction is given by
8
>
>1 uj2 i 6uj1  3uj  2uj1
@u < 2uj jujj 2Dx
for uj > 0;
u
@x >
> 2uj1 3uj  6uj1 uj2
: 2uj  jujj
1
for uj o0:
Dx
143

Higher order interpolation stencils for use in nite-


volume schemes have also been investigated [88].
It is well known that for separated ows where the
velocity changes sign, if upwinding is not used, the
algorithm may not converge. However, in active rotor
codes today, upwinding is used regardless of the Fig. 30. Qualitative sketch of a rotor in hover showing typical
direction of the velocity because of the directionality velocity vector directions.
inherent in the convective terms.

4.2.6. Boundary conditions in CFD calculations


The boundary condition far from the rotor is 4.2.7. Full potential methods
particularly dicult to enforce. The rotation of the The full potential method is a grid-based method that
blades causes air to be drawn into the rotor disk as uses the velocity potential F where ~ u rF as the
shown in Fig. 30. The air coming in through the rotor independent variable. Substituting into the continuity
disk must exit the rotor wake at some point down- equation we have in Cartesian coordinates
stream; elsewhere the air velocity must vanish. Thus, to
limit the size of the rotor domain, many rotor codes use @r @rF @rF @rF
0; 144
a model for the ow downstream often in the form of a @t @x @y @z
uid sink. In addition, the inow velocity above the disk
is often specied from experimental data or modeled. Of where the density is given by the Bernoulli equation
course, it is critical that the models for the upstream and  1=k1
downstream boundary conditions not aect the results k1
r 1 2Ft F2x F2y F2z : 145
close to the rotor. 2

Thus we have a set of two equations in the two


unknowns, density and potential. Here k is the ratio of
Table 1 specic heats. The pressure is obtained from the
Upwind backward stencils for the nth-order formula for the rst equation p Krk where K is a reference constant. The
derivative at the point i: For the forward stencils, the coecients equation for the velocity potential is second order and
are the negatives of the backward and the points are inverted elliptic provided the ow is subsonic and can be solved
n ai5 ai4 ai3 ai2 ai1 ai ai1 ai2 ai3 ai4 b
using standard computational techniques. In transonic
ow as may occur on the advancing side of the rotor,
1 1 1 1 however, the equation is mixed, becoming hyperbolic in
2 1 4 3 2 supersonic regions. Regions of non-zero vorticity are
3 1 6 3 2 6 conned to, at most, the inboard vortex sheets and are
5 2 15 60 20 30 3 60
incorporated by the specication of a jump in potential
7 3 28 126 420 105 252 42 4 420
across the sheet which is determined from the require-
9 4 45 280 840 2520 504 1680 360 60 5 2520
ment of a continuous pressure across the vortex sheet;
452 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

typically the sheet strength is dened by an equation


@g ~
V ave . rg 0; 146
@t
where V ~ave is an average velocity on the sheet.
Linearized versions of Eq. (144), termed small-dis-
turbance equations have been used to predict the local
ow in the transonic region on the advancing side of the
rotor [107]. However, these methods are limited to the
non- or low-lifting case and cannot predict shock
structure. Results for the conservative form of the
equation given above have been obtained by several
researchers [108,109]. Non-conservative forms of
Eq. (144) have also been developed [110112] but these
methods have traditionally been unable to predict shock
location on the advancing side of the rotor accurately.
One advantage of the full potential formulation is that
these computational modules can be coupled to com-
prehensive rotor design codes where trim conditions are
specied within the context of the design process. This
coupling cannot currently be accomplished with Euler or
NavierStokes schemes. On the other hand, this method
cannot be used in the stall region on the retreating side
of the rotor.
Egolf and Sparks [110] solve the steady full potential
Fig. 31. Blade pressure coecient obtained by a full potential
equation in an inner domain using a prescribed wake method [108] at two radial or spanwise locations. (a) r=R 0:5;
specied by a vortex lattice as the far wake boundary (b) r=R 0:96: Here the tip Mach number is MT 0:794; the
condition. Strawn and Caradonna [108] use a conserva- angle of attack is a 121 and the blade is an NACA 0012,
tive formulation of the unsteady full potential equation untwisted rectangular blade. The dash line is the result of Egolf
coupled to a free wake embedded in a comprehensive and Sparks [110] and the dots are from the experiments of
rotor design code. Both hover and forward ight results Caradonna and Tung [113].
are presented and a typical result is depicted in Fig. 31.
Since the far wake is specied from a comprehensive spacings. In this way, no logic is required to locate the
code, the true test of the model is whether the sheets and good agreement is obtained between the
computations predict the blade ow properly. Note the computations and experiment.
generally good agreement with experiment, although
inboard up to about r=R 0:8 the peak pressure is a
little o. Note also the presence of a shock at the outer 4.2.8. Rotorcraft NavierStokes and Euler computations
two locations. Additional results presented in the paper With the improvement in the speed and storage of
indicate that a small-disturbance analysis can reproduce modern supercomputers and the development of parallel
results similar to the full potential code up to an angle of processing, solutions to the NavierStokes equations for
attack of about 81; they note that small-disturbance an isolated rotor are relatively common. In contrast to
analysis could not predict the results for the 121 case. vortex and full potential formulations, the Navier
An attempt to obviate the need for a wake model has Stokes equations contain the physics for vorticity
been discussed by Ramachandran et al. [114] who add a generation at a surface and subsequent convection into
vortical term to the velocity eld according to the wake. In addition, the viscous drag on the blade can
U ~ ~
~O ~V ;
r rf U 147 also be determined for use in computing performance
variables such as the thrust and power coecients
where U ~V is the velocity eld due to the vortical ow. whereas with vortex methods, airfoil lookup tables or
The vortical ow eld dened by U ~V is a model for the some other model is required to calculate drag. Today,
vortex sheets and the tip vortices; it is required to have both Euler and NavierStokes modules often exist in the
the properties that it be non-zero inside the region in same code with the choice set by the user. Two reviews
which the vorticity is non-zero and zero elsewhere. The have been published on the use of these methods;
vortical ow regions move with the local uid velocity Landgrebe [115] outlines the primary contributions of
and the region of non-zero vorticity is constrained NavierStokes and Euler calculations in the United
to have a constant thickness, typically several grid States and Srinivasan and Sankar [116] have discussed
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 453

possible new methodologies. The capabilities of a NavierStokes solutions had also been obtained
European Euler solver has also been presented [117]. [122,123]. Soon after this, full three-dimensional
The computational methods and the implementation NavierStokes solutions were calculated [90]. The
of the boundary conditions vary to a great extent. Thus computed results are compared with the experimental
discussion will focus on the Mach number regime of the data of Caradonna and Tung [113] in the non-lifting,
computation and on the physical results. Typically, transonic regime; results are also produced for a lifting,
results are computed for a single blade with the results NACA-0012 blade. A C-grid is used to describe the
for the other blade(s) assumed to be the same and domain near the blade. The numerical procedure is fully
700,0001,000,000 or more points are used in the implicit in space and the far eld variables are set equal
computations. Most of the work described below uses to zero. Rotor inow conditions are specied using an
a BaldwinLomax turbulence model where turbulence eective angle of attack which incorporates the inuence
has been incorporated. of the wake shed below the rotor which is xed
The rotor is usually started from a given initial wake externally using a rotor design code.
prole and integrated to steady state. This is very Computational results for the blade pressure in hover
dicult, especially in hover and low-speed forward ight from three dierent sources are depicted in Fig. 32.
because of the number of grid points required to resolve Fig. 32(a) shows the Euler results of Agarwal and Deese
the ow for a relatively large number of time steps. [121]; Fig. 32(b) is from Wake and Sankar [90] and
Typically, over two turns of the rotor are required to Fig. 32(c) is from Srinivasan and McCroskey [116]. Note
establish steady state and by this time, the strength of that all of the results show similar behavior and there is
the tip vortex and the inboard sheet is much weaker than no discernible dierence between the solutions at this
that seen in experiments due to numerical diusion. This chord location. The work of Srinivasan and McCroskey
is especially true at the high disk loadings required in [116] uses the thin-layer NavierStokes equations in
helicopters today. The Euler equations are 35 times which only derivatives in the viscous terms normal to the
computationally more expensive than full potential blade surface are retained.
methods with the NavierStokes equations an order of As mentioned earlier all current NavierStokes
magnitude or greater more expensive than Euler calculations of the rotor wake suer from numerical
methods. For this reason, both Euler and NavierStokes diusion in the sense that the vortex system is
calculations of the rotor wake are, at this point in time, considerably smeared. Thus, the fact that the blade
research codes. Euler methods allow the production of pressure distributions depicted in Fig. 32 seem to agree
vorticity near shock waves and can convect this well with experiment is somewhat surprising. However,
vorticity. Provided that the origin of the tip vortex is note that there are signicant relative errors in the
properly modeled, the tip vortex can be convected pressure especially near the tip and at the root and the
properly [118]. Mach number is relatively low. Moreover, it must be
Complicating the computation is the need to trim the pointed out that the results of Fig. 32 depict pressures at
rotor. Many of the NavierStokes computations de- only one section of the rotor blade; these local section
scribed here are, at most, only partially trimmed using errors often lead to large errors in the integrated lift and
an external comprehensive rotor design code. At the normal force coecients and pitching moments which
present time, computational time limitations prevent leads to errors in predicted thrust.
adding a full trim module to the codes. Cyclic and Experimental results become more dicult to predict
collective pitch settings are calculated from the compre- as the tip Mach number increases. In this case, there is a
hensive rotor code and then input to the CFD rotor strong shock which forms and typical results for the
blade code and the ow eld is computed. The thrust blade pressure coecient are depicted in Fig. 33. Here
and power coecients may then be calculated and the the tip Mach number is M 0:877 resulting in a
values reinserted into the rotor design code in which the shock emanating from the blade tip. Note how
rotor is trimmed again. This process may be continued the shock position predicted by the NavierStokes
until the rotor and trim solutions are compatible. computation diers from that predicted by the Euler
Performance parameters can be calculated directly computation.
from the CFD calculations discussed in this section The inuence of wake model has also been analyzed
[119]. Generally, agreement with the model-scale data [91] and a result showing the eect of wake model is
is adequate although signicant dierences of up to depicted in Fig. 34. Improvements in the computational
1020% near the blade tip are observed. Moreover, procedures allow the inboard stations to be computed
moments due to drag eects are more dicult to predict: more accurately and the prescribed wake used in
very little experimental data for drag is available for Srinivasan and McCroskey [116] is no longer necessary.
comparison with the computed results. Note that the peak pressure coecient for the prescribed
Following seminal work [118,120,121], Euler compu- wake boundary condition is signicantly over-predicted
tations have become relatively routine. Axisymmetric whereas the full computation compares well with the
454 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

position is about one blade chord removed from the


position seen in the corresponding experiments.
NavierStokes solutions for the forward ight regime
are relatively rare because these computations are much
more dicult than hover. In this regime, the ow
depends on the age of the wake and is periodic in time.
Further, BVI action may occur (see the next section).
Srinivasan and Baeder [127] have produced solutions for
a forward ight condition for advance ratio m 0:2 and
a blade tip Mach number of Mtip 0:8: The results at
rotor phase angle c 1201 are not signicantly dierent
from those depicted in Fig. 34 except that the shock
moves farther down the airfoil to x=c 0:6:
Comparisons between ight test data, a prescribed
wake analysis and a full NavierStokes result have been
made by Rajagopalan and Mathur [128] and a typical
result is depicted in Fig. 35. Note the deviation from
the test data; this comparison is typical of current
capabilities.
The results discussed above for the NavierStokes
equations include a time-stepping algorithm; thus
depending on the accuracy of the time-stepping proce-
dure, the discretization error can grow rapidly with time.
This means that the accuracy of blade loads is
substantially degraded and this is exacerbated at full
scale by the aeroelastic deformations and apping
motions. In Fig. 36 are results from Bangalore and
Sankar [88] for a UH60A rotor for two values of rotor
phase angle. Note the signicant discrepancy between
the computations and experiment on the retreating side
while the early time results are fairly accurate.
Ahmad and Duque [87] include moving embedded
grids to calculate the solution for the rotor system of the
AH-1G helicopter which utilizes a two-bladed rotor.
The authors note that the embedded grid procedure
allows a more ecient and more accurate calculation of
the ow near a rotor blade undergoing dynamic
motions; the rotor is partially trimmed externally. The
calculation is started from freestream conditions and ve
complete rotor revolutions are calculated. A typical
result for the rotor wake is shown in Fig. 37. The wake
streaklines exhibit periodicity in about two rotor
revolutions; note that while the streakline patterns
clearly show the trajectory of the tip vortex, it is seen
Fig. 32. Blade pressure coecient for a lifting rotor in hover. that particles are diusing as they interact with the
Mtip 0:44; collective pitch 81; Re 106 ; the dots are the following blade. This is an indication that the magnitude
experimental data of Caradonna and Tung [113]. In this gure, of the vorticity in the tip vortex may be very small. The
y denotes the radial or spanwise location. (a) From Agarwal blade pressure and section normal force show signicant
and Deese using the Euler equations [121]. (b) From Wake and dierences when compared to experiment and the power
Sankar [90] using NavierStokes. (c) From Srinivasan and is overpredicted by 15%. To get a feel for the size of the
McCroskey [124] for thin-layer NavierStokes; the solid line is calculation, the complete unsteady calculation takes a
the computation.
total of 45 h of single processor CPU time on a Cray C-
90 supercomputer and generates 40 Gb of oweld data.
experimental data. Wake and Baeder [126] are able to One method designed to oset the numerical diusion
resolve the interaction of the tip vortex shed from the process is described in a review by Steinho [129].
preceding blade with the following blade; however, its Steinho [129,130] prevents the diusion of vorticity by
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 455

Fig. 33. Blade pressure coecient for tip Mach number M 0:877 in hover at two radial or spanwise locations [125]; blade angle 81
and Re 3:93 106 : Here the solid line is the NavierStokes computation, the dash line is the Euler computation and the dots are the
results of the experiments of Caradonna and Tung [113].

Fig. 35. Blade loads for an isolated two-bladed rotor in


forward ight; m 0:22 and two blade chord locations [128].

inserting an external force in the NavierStokes


equations acting in a direction normal to the vortex
sheet and one such choice is
F~E en# o
~; 148
where e is a parameter which controls the size of the
convecting regions of signicant vorticity (i.e. the
inboard vortex sheet and the tip vortex), and n# is a unit
Fig. 34. Blade pressure coecient for the parameters of Fig. 33 vector in the direction of the normal to the boundary
[91] at two radial or spanwise locations. (a) r=R 0:5; (b) of the non-zero vorticity region. Application of this
r=R 0:89: The solid line is the computation with no wake concept to the axisymmetric ow eld induced by a
model, the dash line uses a prescribed wake [124] and the dots Lamb vortex results in the existence of a steady-state
are the experiments of Caradonna and Tung [113]. solution for the velocity eld and hence a steady-state
456 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

solution for the vorticity; that is, there is no diusion of


vorticity. Thus, Steinho refers to this method as
Connement. A typical result is depicted in Fig. 38
which shows a cross section of a rotor for the parameters
of the Caradonna and Tung experiments [113]. Note
that the vortex is well preserved at least to one turn of
the rotor.
A more traditional method to increase accuracy of
calculations is to use high-order dierencing schemes as
discussed in Section 4.2.5 [103,105,106]. A typical result
from [106] for a nite-dierence scheme is shown in
Fig. 39. Note that the swirl velocity prediction for the
ninth-order scheme preserves the amplitude of the tip
vortex fairly well. They note that in addition to using
higher order methods it is necessary to align the grid
with the principle ow direction. For all problems
considered, the ninth-order scheme is considerably more
accurate than a fth-order scheme. High-order methods
for nite-volume schemes have been reported [103] and
high-order accuracy schemes with grid adaptation have
also been used [131]. The use of unstructured grids for
this purpose has also been discussed [89].

4.3. Blade aeroelasticity and trim

Complicating the wake calculation is the fact that the


rotor blades undergo complicated motions in order to
Fig. 36. Blade pressure coecient for Mtip 0:628; m 0:3; for balance moments and because of their aeroelastic
the UH-60A baseline rotor at (a) y=R 0:775; c 301: (b) On
properties. As has been mentioned, a rolling moment
the retreating side at y=R 0:4; c 3201 [88]. The experi-
mental data are from Lorber [144].
imbalance exists in forward ight and so the rotor must
be trimmed; that is, a balance of moments must be
achieved. This balance is obtained by pitching the angle
of attack as the rotor negotiates the cycle. In addition,
the rotor blade may undergo rigid ap, and leadlag
motions in order to balance moments in the other
coordinate directions (Figs. 3 and 40). These three rigid
body motions are possible only for articulated blades.
Hingeless rotors may undergo rigid body pitch.
In addition, depending on how the blade is attached
to the hub, aeroelastic properties of the blades may
become important; they bend both in-plane and out-of-
plane and undergo torsional motions of perhaps large
amplitude. In addition, all of these motions may be
coupled. For bearingless rotors, aeroelastic eects
dominate. In this section we give a brief outline of the
nature of the aeroelastic phenomena and of the
equations governing rigid body motions. We consider
in this section only isotropic materials although compo-
site materials have been considered [132] and a beam
theory for anisotropic materials has been presented [133].
Consideration of instability of these motions is a
necessary ingredient to the design of rotor blades.
However, space limitations preclude a discussion of this
phenomenon here. There are many dierent types
Fig. 37. Streaklines for a rotor wake at m 0:19: Flow of analyses and this section is meant to give only
periodicity is established in about two rotor revolutions [87]. a brief overview of representative equations and
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 457

Fig. 38. Tip vortex evolution: (a) standard CFD computation; (b) with connement [129,130].

Fig. 39. Tip-vortex velocity eld one chord length behind a


model rotary wing operated as a xed wing for several dierent
high-order discretization schemes [105,106].

methodologies. Additional details may be found in


helicopter dynamics books [20,134136] and papers
[137141]. A comprehensive review [142] and a summary
of the types of engineering beam theories has also been
published [143]. Of course this reference list is not
complete and is intended to be merely a starting point
for analysis of the subject.
The aeroelastic and trim problems introduce addi- Fig. 40. Top: a sketch of the trim motions: ap, leadlag and
pitch of a single helicopter blade. From Peters [138]. Bottom:
tional partial and ordinary dierential equations which,
aeroelastic modes: ap bending, leadlag bending, torsion.
in the most general case, must be solved simultaneously
with the NavierStokes equations, a vortex model for
the rotor wake, or some other wake model. In general, 4.3.1. Rigid blade motions
computer time requirements to fully couple the aero- For an articulated rotor the rigid body dynamics of
elastic and trim equations with the NavierStokes the rotor usually dominate aeroelastic eects. In the
equations or a complex free wake model is prohibitive most general case, the blade will undergo three types of
and in these cases trim is only performed a xed number rigid motions in order to balance moments:
of times per cycle; often the rotor is trimmed on the * out-of-plane apping;
order of once every 10201 of rotor motion. Trim and * in-plane leadlag; and
aeroelastic analyses most often use simpler models for * pitching: oscillating angle of attack.
the rotor wake such as rigid or prescribed wakes, and
dynamic inow [77,78]. A similar comment applies to Thus the three variables to solve for the dynamic motion
the aeroelastic calculation. of the blade are the leadlag angle, the apping angle
458 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

and the blade pitch angle. These angles are obtained by where K is a leadlag hinge spring. Note that for small
requiring a moment balance in each of the directions leadlag angles, the force Fx is the total drag force
dened by the angles described above. Consider rst the composed of the viscous and induced drag. A full
case of a constant chord, rigid blade with no ap hinge derivation of the non-linear versions of apping and the
as depicted in Fig. 40. We also ignore the presence of a leadlag equations has been presented [138,140].
hinge oset. From the gure, it is apparent that the The pitch angle equation is similar to that of ap and
blade is subject to two types of acceleration which can be leadlag, and is given by, in the absence of ap or lead
decomposed to balance the aerodynamic lift: these are lag coupling,
the centrifugal acceleration and the acceleration due to
y. ly y gf MP o2 yi ; 153
the time dependence of the apping angle. To balance
moments we take r * the linear momentum equation where MP is the applied moment about the feathering
and integrate over the span of the wing and for small axis of the blade. Here ly 1 o2 ; where o2 Ky
ap angle the linearized equation governing ap angle is If O2 ; and If is the total integrated moment of inertia
given by about the feathering axis. yi is the control pitch angle.
Z R Z R Also gf is the Lock number based on If :
mr * b. mO2 r * r * b dr * Fz r * dr * ; 149 It should be emphasized that the equations presented
0 0
here are the simplest possible for a rotor system. In
where in this linearized case, Fz is the lift; here m is the many cases ap, leadlag and pitch motions are coupled.
mass of the rotor blade. We note that the moment of Analytical formulas for these angles in the case of auto-
inertia about the ap hinge is given by rotation have been presented [139].
Z R
IB mr * 2 dr *
0 4.3.2. Blade aeroelastic deformation
For a hingeless blade which is attached to the hub by
so that we may write the equation as
Z a cantilever, aeroelastic phenomena usually dominate
O2 R
b. b r * Fz dr * ; 150 rigid body motions. Rotor blades undergo three modes
IB 0 of aeroelastic motions:
where r * is the dimensional radial variable. If we non- * out-of-plane bending or ap bending;
dimensionalize Eq. (150) and linearize * in-plane bending or lag bending; and
Z
g 1 * twist or torsion.
b. b rL dr; 151
2 0
These motions are depicted in Fig. 40(b) and are
where g rcmR4 =IB is the Lock number and is the ratio responses of the blade to induced forces.
of the aerodynamic forces to the solid inertia forces. The A rotor blade is viewed as a rotating beam with a
Lock number gB5210 for most rotor blades. Eq. (151), exural rigidity and the equation which governs the
is recognized as a driven springmass system, with amplitude of out-of-blade bending may be derived from
frequency of 1=rev: Eq. (151) is usually solved by modal classical beam theory as [134,135]
decomposition since the equation is linear.  Z R 0
To complete the calculation of the apping angle, the mz. EIz00  O2 z0 mr * dr * FOB r * ; t; 154
lift force L Fz must be calculated. Models used to r*
calculate the lift range from momentum theory and
where E is the modulus of elasticity and the prime
blade element theory to a full solution of the Navier
denotes dierentiation with respect to r * : In this
Stokes equations. For momentum theory, in forward
linearized theory the force FOB r * ; t is approximately
ight the lift coecient can be decomposed into
the lift. This is a fourth-order, linear partial dierential
aerodynamic coecients due to blade pitch, twist,
equation and is second order in time. The usual method
inow, apping velocity and ap displacements from
of solving this equation is to use the separation of
which analytical formulas for the required angles (there
variables technique to generate modal equations having
is pitch-ap coupling in this case) may be obtained ([20],
a form similar to the ordinary dierential equations
p. 187).
given above for the rigid blade motions.
In a fully articulated rotor, leadlag motions are
The in-plane bending motions are governed by a
coupled to the apping motion, and using the same
similar equation but including the Coriolis force and is
reasoning as given above for the apping motion the
given by
linearized governing equation for the leadlag angle is of
 Z R 0
the form
Z 1 mz. EIz00  O2 z0 mr * dr * O2 my FIB r * ; t:
r*
z. Kz 2bb g rFx dr; 152
0 155
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 459

In this linearized case the force FIB r * ; t is approxi- aeromechanics; however because of the integral role
mately the total drag, given by the sum of the viscous that experiments play in uncovering the behavior of
drag and the induced drag. rotor wakes, a paper such as this would not be complete
Both of these aeroelastic equations are subject to four without a section on experiments.
boundary conditions which are standard in beam As has been pointed out, the study of rotor wakes has
theory. Note that four boundary conditions are required employed various methodologies in an eort to develop
and three of these are an accurate model of the ow both near the blade and in
@2 z @3 z the ultimate wake. However, there is no guarantee that a
zrr* ; t 2
R; t R; t 0: 156 given model will accurately reect reality. Thus experi-
@r * @r * 3
ments are required to validate models of the rotor wake.
The rst of these conditions says that the blade has zero Moreover, because of the diculty in measuring the
deection at the root, while at the tip, the bending entire wake and blade ow simultaneously, it has been
moment and the shear both vanish. For an articulated common practice for appropriately targeted experiments
blade the bending moment is zero at the root or to guide the development of a model: historically there
@2 z * has been an intimate link between experiment and
rr ; t 0 157 analytical/numerical work within the rotorcraft com-
@r * 2
munity. Moreover, experimental measurements have
and for a hingeless blade, which is connected by a
traditionally been the foundation of helicopter design
cantilever, the root slope (i.e., coning angle) is taken to
and in this section we examine the various experimental
be specied:
approaches which have been used in the past, as well as
@z * those that are currently being used.
rr ; t bco : 158
@r * Unlike the xed-wing environment, measurements of
In addition to physical twist, a rotor blade is subject the ow over a rotor blade are complicated by the rapid
to stresses which result in aeroelastic twist or torsion. motion of the blades. The ow eld thus contains air
The equation governing torsion is more complicated velocities which range from transonic to supersonic near
than the equations for aeroelastic lag and ap bending the blade tips to essentially incompressible a short
and the linearized equation governing torsion is ([20], distance into the wake. Thus, the methods used in a
p. 416; 135, p. 96) given measurement depend critically on the objective of
 Z R 0 the experiment.
Iy y.  GJ y0 0 Iy mO2  mxI z. r * z0 mO2 xI dr * The interplay between experiment and modeling
r* seemed to begin in 1950s when ight test experiments
Mapp ; 159 were used to validate wake models based on momentum
theory. These measurements were often done at low disk
where the prime indicates dierentiation with respect to
loadings and in hover or low-speed forward ight under
r * ; GJ is the primary torsional stiness, xI is the
nominally steady ow conditions [3]. Landgrebe [52]
moment arm about the feathering axis. In the limit
visualized the wake of a two-bladed rotor with smoke
GJ-0; this equation reduces to coupled rigid pitch and
emitted from rakes in a single plane. The results in
ap. This is a second-order equation in space and is
Fig. 41(a) depict the tip vortex and the inboard vortex
subject to the boundary conditions
sheet in a two-dimensional cross section of the wake.
yrr* ; t 0 160 Much of the prescribed wake methodology was devel-
and oped from analysis of photographs such as in Fig. 41(a)
to determine vortex trajectories for use in developing
@y useful computational models.
R; t 0: 161
@r *
The rst of these boundary conditions corresponds to a 5.2. Levels of the experiments
xed end at the root with the second corresponding to a
free end at the tip. Linearity allows solution by modal Experiments have been performed on essentially two
decomposition. distinct levels although intermediate variations do exist.
These levels are model-scale experiments in which the
underlying physics of the ow can be more easily
5. Experimental methods identied and full-scale experiments which typically can
include the complex blade motions required to trim the
5.1. Introduction rotor. Model-scale experiments are often performed
using rigid blades. In order to preserve adequate
The experimental investigation of rotor wakes chordwise resolution and to produce the interaction of
would not normally fall under the classication of strong vortices with surfaces, model rotors generally
460 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

inertia and the expense of optical systems. Blade-


mounted instrumentation such as sensors of surface
pressure, acceleration and shear stress must perform
under extreme radial acceleration, and still have
sucient signal-to-noise ratio when transferred from
the rotating to the xed co-ordinate system. These
considerations limit the applicability of many diagnostic
advances demonstrated in xed-wing aerodynamics and
bench-scale water tunnels.

5.3. Data types

There are roughly four dierent sets of aerodynamic


data which are sought. These sets of data are (1) ow
visualization of the rotor wake by smoke or some other
non-intrusive means such as laser Doppler velocimetry
(LDV), wide-eld shadowgraphy, Schlerien techniques
or particle image velocimetry (PIV); (2) direct measure-
ment of the blade (or airframe) pressure distribution; (3)
measurement of rotor inow velocities. Experimentally
generated rotor inow conditions are often used as
boundary conditions for CFD studies; and (4) direct
measurement of the wake velocity eld. Methods for
acquisition of this data range from intrusive methods
such as hot wire probes to nominally non-intrusive
means such as those mentioned above. The ow in the
rotor wake can be visualized using smoke and other
seeding materials in conjunction with a laser light sheet
Fig. 41. Experimental results from Landgrebe [52]. (a) A as well. Lorber [144] has described the use of several
sample smoke visualization of a slice of the rotor wake for a dierent visualization and measurement tools for the
two-bladed rotor. (b) Tip-vortex trajectory for a six-bladed comprehensive description of helicopter ow elds.
rotor for several tip Mach numbers. LDV and PIV are advanced measurements techniques
which are often used to measure the three-dimensional
have diameters of at least 1 m: Rotational speeds in these ow eld both in and around the tip vortex and in the
model-scale experiments usually range from 1000 to balance of the wake. Both methods use neutrally
2500 rpm: Because the air ow far from the rotor should buoyant particles to measure the local uid velocity.
be negligible, the facility walls must be at least 1 m away LDV uses a pair of light beams to measure a single
from the hub. The blades must be strong enough to velocity component by measuring the Doppler shift each
withstand the rapid rotation and so bullet-proof time a particle crosses the volume dened by the
windows are essential parts of a rotorcraft ow crossing of the two light beams. PIV employs a laser
measurement. light sheet to illuminate a plane within the ow which
Full-scale rotor blades are often several meters long can be photographed at given time intervals yielding two
with a span to chord ratio typically ranging from 5 to 20 of the three velocity components in a given region of the
depending on the design. A typical full-scale rotor speed ow. The third component must be obtained from
is about 300 rpm with tip speeds designed to achieve photographs taken in the plane normal to the given
Mach MaB0:6: Full-scale experiments are generally plane just described or from the use of multiple cameras.
trimmed and so the rotor blades ap, teeter, pitch, bend Both of these methods, with improvements in computer
and twist; these motions pose problems for mechanical capabilities and technical improvements in laser equip-
probes and it is sometimes dicult to locate a point on a ment and digital image processing can now be viewed as
rotor blade in time and space with the accuracy needed routine, although improvements continue to be made.
for a pressure measurement. Because of the wide range
of velocities, various measurement techniques may not 5.4. Experimental results
be appropriate locally; for example, where the wake is
incompressible, density-gradient visualization techni- The smoke visualizations shown in Fig. 41(a) were
ques will provide little information. Light-scattering used to develop analytical expressions for the time-
techniques encounter the tradeo between seed particle averaged radial and axial positions of the helicopter
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 461

wake. The axial coordinate of the tip vortex could be addition, measurements have often been conducted on
accurately described by a formula of the form rotorbody interactions as well as for isolated rotors.
2p Wide-eld shadowgraphy has been used to investigate
z% k1 cw for 0pcw p ; the wake of a tail rotor [145]. This technique has the
b
major advantage that it is non-intrusive and does not
  require seeding. Instead density gradients are employed
2p 2p
z% z%cw 2p=b k2 cw  for cw X ; 162 to locate a tip vortex and combined with pulsed lighting
b b
and video, enables quantitative analysis of the projec-
where z% is a non-dimensional distance below the rotor tion of vortex trajectories. Using the same method,
disk and is normalized on the rotor radius. Here cw is measurements of the tip-vortex geometry in the wakes of
the azimuthal wake coordinate relative to the blade, and two main-rotor models in hover have also been reported
b is the number of blades. k1 and k2 are parameters [146]. Some scatter in the data due to rotor wake
which are obtained from the experiment. The radial instabilities was reported and a short-wavelength
coordinate of the tip vortex was found to correlate well instability of the vortex core was also observed. The
with an equation of the form wake of a model scale V-22 rotor test in hover has also
r% A 1  Aelw cw ; 163 been investigated [147]. Improvements in image quality
by including the addition of a beam splitter and a
where A 0:78 and lw 0:145 27CT ; r% is the radial reorientation of the camera system have also been
coordinate of the tip vortex normalized on the rotor reported [148]. An example of the visualization of a tip
radius. Similar wake coordinates are given for the vortex using wide-eld shadowgraphy for visualization
inboard sheet. These formulas have been periodically of the tip vortex is depicted in Fig. 42 [81].
updated for dierent conditions and as additional Interactions between tip vortices in the wake have
experimental measurements become available [6,46,53]. been seen in studies as noted previously [45,5254].
These simple wake coordinate formulas are used in Recent experiments [85] have shown these types of
prescribed wake models to predict the rotor airloads. interactions clearly in low-speed climb and typical
An extensive data set on which many theoretical and results from these experiments are shown in Fig. 43.
computational schemes are calibrated is derived from Vortices 3 and 4 are from dierent blades and note that
the experiments of Caradonna and Tung [113]. The they roll over each other approximately one rotor radius
experiments were conducted in hover on a two-bladed from the rotor disk. The time scale for the entire process
untapered and untwisted blade of aspect ratio six with is about one and a half rotor cycles or about 90 Hz for a
an NACA0012 airfoil section. They measured blade two-bladed rotor. These interactions would be expected
pressure distribution and wake geometry; the pressure to be more severe and occur more often during a rotor
distribution was measured using three pressure transdu- cycle in descent [81].
cers along each blade span. The wake properties were Laser light sheet methods are useful for determining
measured with a hot-wire probe oriented with the wire vortex location; however, more sophisticated methods
being tangent to the tip-path plane. Several rotor speeds are required for measuring the three-dimensional
were tested and the two-bladed rotor was substantially velocity eld and vortex core size. In addition to
rigid. Many of the papers cited previously have
validated wake models based on comparison with this
data set.
In the Caradonna and Tung experiments, the tip-
vortex trajectory resembles the results presented by
Langrebe [52] and the tip-vortex strength was found to
be close to the maximum bound circulation on the
blade. In addition, the vortex structure was found to
closely resemble the classical Rankine vortex. However,
the measured and computed blade loading using a
lifting-surface approach with a prescribed wake do not
match, leading to the conclusion that the blade spanwise
loading cannot be predicted accurately without a better
model for the vortex location and strength.
In the last 20 years there has been a signicant
improvement in experimental techniques as various Fig. 42. Visualization of the tip vortex shed from a two-bladed
types of non-intrusive measurement techniques have propeller using wide-eld shadowgraphy [81]. Note that the
replaced intrusive procedures such as hot lm and probe vortex is not diused and irregular vortex motions occur farther
methods and the use of passive scalar techniques. In down in the wake similar to those shown in Fig. 24.
462 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Blade bound circulation and the ow in a vortex have


been investigated by a number of researchers [153158].
Mba et al. [154] measured rotor blade circulation in
hover by contour integration of laser velocimetry data.
Thompson et al. [153] used submicron incense smoke
particles to measure the ow eld in the core of the tip
vortex in hover. The core axial and tangential velocity
proles at two thrust coecients show evidence of
discrete shear layers rolling up to form the vortex
core. They also constructed vortex trajectories in the
near wake of a single-bladed rotor using still photo-
graphs taken from mineral oil seeding, frozen using an
argon ion laser sheet strobed using a synchronous
chopper.
Measurements of an entire planar air velocity eld in
rotor ows were performed by Funk et al. [159] using a
double-camera system and spatial correlation veloci-
metry. The technique is successful in the wake, and a
periodic variation was constructed from video image
pairs obtained at several azimuths. However, the spatial
resolution is as yet inadequate to measure velocity inside
the tip-vortex core. Time-averaged velocity elds have
been measured using Doppler global velocimetry [161].
The technique was applied to measure time-averaged
velocity near the empennage of a small-scale helicopter
model, and compared to laser velocimeter data. This
technique has the advantage that every pixel location on
the digitized image (640 480 resolutions are common
now), gives one velocity vector. The disadvantages are
the need for multiple cameras, the huge data reduction
task, involving several gigabytes of data, and the low
signal-to-noise ratio at low speeds.
In a relatively comprehensive set of experiments,
Fig. 43. Sequence of vortex roll-up from the individual blades McAlister et al. [157] have measured the velocity eld
from the experiments of Caradonna and Komerath [85]. Here, behind a blade tip using three-color laser velocimetry
the collective pitch is 91 and the climb rate is 3:5 ft=s: The ow and a typical result is presented in Fig. 44. Note that the
is from right to left. velocity Vz is typical of a vortex and there is signicant
axial ow along the vortex axis (the maximum value
of Vx is about 20 m=s) and directed back toward the
resolving problems with intrusive methods LDV has wing.
been employed to measure the three-dimensional velo- A similar set of experiments has been conducted by
city eld, at the steep price of having to use particle Leishman et al. [46]. The one-bladed rotor system
seeding and particle velocity as an indicator of ow consists of a rectangular blade with an NACA 2415
velocity [149]. Desopper et al. [150] discuss incense sectional prole. Measurements were made at low tip
seeding to measure wake velocities over a wide range of speed and at a Reynolds number ReB105 for the rotor
tip speeds. Liou et al. [151] measured the unsteady in hover. Results for the velocity eld within the vortex
velocity eld around a rotor blade with controlled pitch and circulation were presented and compared with
excitation of 11 amplitude, and correlated them with models for the vortex core as discussed previously.
blade surface pressures and unsteady inow prediction PIV has also been used to measure the three-
codes. Laser schlieren visualization of the compressible dimensional velocity eld in the wake. As mentioned
ow was used with tomographic reconstruction to earlier in this section, PIV uses neutrally buoyant
obtain fully three-dimensional features of the rotor seeding particles and a pulsed light sheet to measure
wake [152]. the velocity eld. If the time between pulses is known,
These advanced techniques are useful for measuring the velocity of each particle can be measured and a series
the circulation and structure of the tip vortex on a rotary of measurements can be used to map out the ow eld.
wing is crucial for the development of wake models. Heineck et al. [160] have used this technique to measure
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 463

Fig. 44. The three-dimensional velocity eld downstream of a rotor blade along a horizontal sweep across the trailing vortex. The time
frame shown corresponds to the motion of the rotor blade from 0.3 to 1.0 chord lengths past the measurement location as noted in the
sketch. The rotor speed is 1100 rpm and the blade tip is at y 0 [157].

the velocity eld within the core of a tip vortex for a As an example of the fact that a combination of
two-bladed rotor in hover up to a vortex age of 2701: methods is required for a comprehensive understanding
Flow visualization combined with a pulsed copper of the physics of the rotor wake, a measured rotor inow
vapor laser sheet and laser velocimetry was used to show velocity eld is commonly used as a boundary condition
that the edge of the inboard vortex sheet of a rigid and in both vortex method algorithms and in CFD analyses.
untwisted, two-bladed rotor rolled up into a discrete A survey of models for non-uniform rotor inow for use
structure with circulation strength approximately half in both momentum approaches and in detailed CFD
that of the tip vortex, and opposite in sign [155]. This analyses of rotor wakes has been presented [162];
eect has also been observed in video images [156], but additional experimental and computational results have
this phenomenon has normally not been included in been given by Elliott and Altho [163], Hoad et al. [164]
rotor wake models. and Hoad [165].
464 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

6. Special topics

6.1. Bladevortex interactions

A major impediment in the development of quieter


helicopters is the physical phenomenon known as blade
vortex interaction (BVI). Unlike the wake of a xed
wing which rapidly moves away from its generating
surface, the rotor wake shed from one blade may
interact signicantly with the other blade. This phenom-
enon is a major source of noise generated by the
helicopter. BVI is most severe when the vortex
approaches the blade approximately aligned with the
spanwise axis of the blade. This means that the
interaction between the vortex and the rotor blade is
nearly two dimensional. BVI is rare in hover, can occur
in forward ight, but may be particularly severe during
maneuvers, in vertical descent and in landing (forward
ight descent). There is a large body of work in the
literature on various aspects of BVI and the noise
produced as a result.
It is generally acknowledged that three types of
interactions between the rotor blade and the tip vortex
are most common. These interactions are

* almost parallel interactions (Fig. 45(a));


* almost perpendicular interactions (Fig. 45(b)); and
* direct collisions in which the core of the vortex is
locally destroyed [166]; (Fig. 45(c)).

BVI noise is emitted as a result of the highly unsteady


interaction of the vortex with the blade which involves a
rapid acceleration and possible distortion of the vortex
which is easily demonstrated in a simple two-dimen-
sional problem [167169]. This occurs as the vortex
passes above or below the surface of the blade and
continues until the vortex passes the trailing edge. The
BVI problem is considerably more complex because, in
general, the structure of the vortex is unknown and must
be calculated, the time scale of the interaction can be
much shorter than the time scale associated with the
evolution of the wake, and a given vortex may interact
with several following blades and other tip vortices
[144,170172].
BVI noise is one of the two major components of
impulsive noise associated with the ow past the rotor
blades, the other being high-speed impulsive noise due to
the high tip Mach number on the advancing side of the
rotor [173,174]. BVI can occur on both the advancing
and retreating blade sides, but from an acoustic point of
Fig. 45. Sketch of the most common BVI events: (a) almost
view, the interactions on the advancing side are more
parallel interactions; (b) almost perpendicular interactions; and
important because of the higher Mach number there. As
(c) oblique collision. A sketch of the tip vortex shed from the
pointed out by McCroskey [5] and Yu [175] and which interacting blade is also shown.
can be veried by a simple dimensional analysis, the
most important parameters in the bladevortex encoun-
ter are the strength of the vortex and its distance from
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 465

the blade. Prediction of the magnitude of the acoustic of authors, both explicitly [176] and implicitly by use
signal in the far acoustic eld is limited by the accuracy of a standard rotor design code which uses an
of the calculation of the location and strength of the tip inviscid-based approach to calculate the wake.
vortex; small changes in miss-distance can result in * The blade pressure, which must be input into an
signicant dierences in the nature of the acoustic eld. acoustics code, depends strongly on the local
Moreover, BVI calculations are generally made with trajectory of the vortex and must be calculated very
a coupling of a CFD or some other rotor wake code to accurately [178,178].
a comprehensive rotor design code in which the trim * Reasonably good predictions of blade pressure for a
conditions and aeroelastic properties of the rotor blades single vortex in the parallel case can be made if the
are prescribed. While it is generally acknowledged that structure of the vortex is specied; the predictions are
BVI is a local event, a comprehensive rotor design code not, in general, accurate if the structure of the vortex
requires resolution of the entire rotor wake. This means must be calculated dynamically [172,176,179]. How-
that a single aerodynamic module must accurately ever, when the structure of the vortex is calculated
represent in both space and time, the generation of the along with the balance of the solution of the rotor
tip vortex, its motion, its interaction with one or more wake, the results are not nearly as good.
blades, and interactions with other tip vortices and the * Multiple interactions occur during a typical rotor
inboard sheet. Indeed, wake calculations using a cycle with the most intense interactions taking place
comprehensive rotor code are limited to an accuracy on the advancing side of the rotor and just inboard of
of about 10201 in rotor phase angle. Consider the the rotor tip [144]. However, the interactions on the
almost-parallel situation and suppose the rotor speed is retreating side can be intense as well.
300 rpm and an advance ratio is 0:15: At room
temperature this corresponds to a tip Mach number A detailed study of multiple self-generated BVI
MB0:25 for a rotor radius of 3 m and a forward ight encounters has been given by Hassan et al. [176]. They
speed of about 15 m=s 31 mi=h). The time scale of solve the Euler equations to produce results for a
passage of the vortex over or under a blade will vary number of BVI encounters during a single rotor
over the rotor span but near the tip on the advancing revolution. The solutions are found to be extremely
side, the time scale for passage past a blade having sensitive to the value of the vortex core radius chosen.
a chord of about c 120 mm is A typical result at two inboard stations is depicted in
Fig. 46. In these results, BVI commences at around
c 120 m cb B601: Note that agreement with the experimental
tB B 1:5 103 s;
W 81 m=s values is good until about cb 901; note the dierences
between vortices of dierent core radius.
where W is the tip speed. On the other hand, at this
Lorber [144] has experimentally investigated BVI in
rotational speed, c 101 corresponds to a time of 5
the forward ightdescent regime for a four-bladed
104 s and so this rotor phase increment is of the same
rotor and his results for advance ratio m 0:176 are
order as the overall time scale of the interaction. Thus it
depicted in Fig. 47. Note that most of the BVI events
is not surprising that analyses of the BVI phenomena
occur on the advancing side of the rotor; these events are
which are modeled using classical rotor design codes can
seen to occur from a radius of 0:4R all the way out to the
likely miss many of the short-time-scale phenomena
tip. Parallel interactions on the advancing side are
which occur during BVI events. The situation is even
concentrated around the wake azimuth cb 40501:
more acute if the local length scale of the ow, which is
The retreating side events are seen to occur farther
the vortex core radius and not the chord of the airfoil, is
outboard and between cb 290 and 3201:
used in the scaling. Indeed in order to match experi-
The strength of the tip vortex as measured by its
mental data, Hassan et al. [176] had to use a much
circulation is a strong function of the blade tip shape
weaker vortex than that indicated by experiments in
from which it was shed. Consequently, much research
their computations using a comprehensive rotor code for
has been done to determine if BVI noise can be reduced
trim and an Euler solver and discrepancies were still
by suitable design of the tip shape [171,180].
present. They attribute some of the discrepancies to the
assumed invariance of the vortex core structure.
Despite the complexity of the problem, past work has 6.2. Interactional aerodynamics
shown that several fundamental characteristics seem to
be common to VBIs and some of these are In recent years there has been a continuing eort to
manufacture cheaper helicopters which provide higher
* The essential physics of the process appear to be performance in terms of payload capability and forward
captured by inviscid ow theory. This can be shown speed and maneuverability. This trend requires that the
on the basis of a dimensional analysis of the structure rotor attain higher tip speeds and that the rotor disk
of the vortex [166] and has been noted by a number plane be closer to the airframe; the higher rotor speeds
466 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Fig. 47. Mapping of BVI events on a four-bladed rotor: (a)


Section thrust at m 0:176 as a function of span. The low-
frequency content has been ltered leaving the high-frequency
BVI content. (b) Mapping in the rotor disk for the same
advance ratio. Size of the circles indicates the magnitude of the
event [144].

developed [181] and these interactions are summarized


Fig. 46. A comparison of predicted and measured surface in Fig. 48. Two of the most important interactions are
pressures for the model rotor under BVI conditions at two main-rotortail-rotor interactions and wakefuselage
dierent radial or spanwise locations for two values of the interactions. The relative importance of these interac-
vortex core radius. Three BVI interactions occur during this tions depends on the relative proximity of individual
time range. Here RBAR y=R and rv=C is a dimensionless components, their size, their shape and ight conditions
vortex core radius. The tip Mach number M 0:67; advance (e.g. low speed, in ground eect, ight with sideslip).
ratio 0.1632. The distributions are at 3% chord [176]. Here The function of the tail rotor is to balance the torque
azimuth refers to rotor phase angle.
of the main rotor, thus providing better control in hover
and low-speed forward ight [5]. Performance may be
required generally results in higher disk loading. Thus, adversely aected if the main rotor wake impinges on
the wake shed by the rotor may impinge on one or more the tail rotor. The diculties in designing the empennage
of the other structures on a helicopter. In the last 10 and tail rotor due to poorly understood interaction
years, helicopter aeromechanics has advanced to the eects have been discussed [182]. In addition this
point where these interactional aerodynamic phenomena interaction can be a signicant source of noise [5].
have received considerable attention. A detailed classi- The vortex wakefuselage interaction has perhaps
cation of a variety of aerodynamic interactions received the most attention within the last 10 years or so.
between major components of the helicopter has been This interaction process is especially intense in hover,
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 467

Fig. 48. Summary of the many interactions between the major components of a helicopter [181].

climb and low-speed forward ight since the wake is tip vortex and its induced pressure distribution existed.
transported almost vertically downward. The impinge- During this time period attempts to include airframe
ment of the wake on the fuselage may aect the handling eects on the wake included various levels of empiricism
qualities of the aircraft because additional periodic and beginning with the use of a prescribed wake structure
even impulsive loads on the fuselage will occur. While [187,188]. One approach used the prescribed wake
the time frame over which these additional loads may be structure to specify the trajectory of the tip vortex as it
small, the amplitudes of the loads may be large and so approaches the airframe; a xed oset distance denes
fatigue may be hastened. Fuselage loads can also be the limiting distance allowed between the tip vortex and
transmitted to the cabin in the form of a low-frequency the airframe [188]. The results show fair agreement with
vibration of the fuselage shell and thus can be annoying experiment; the time-averaged mean pressure on the top
to passengers. of the airframe is in general underpredicted. Similar
The very rst work on the wakefuselage interaction discrepancies appear in the instantaneous pressure on
[183185] established that a smaller clearance between the top of the airframe. Additional work in this area
the rotor and the airframe will lead to higher induced shows similar results [189,190].
airloads and vibration levels on the airframe. It is well Beginning in the 1990s, the tip-vortex trajectory and
known that the main feature of the pressure distribution the velocity near the tip vortex began to be measured
due to a vortex moving above a surface is a local suction [191]; results for the pressure distribution on the surface
peak of an amplitude that depends on the vortex of the airframe were also presented [192]. This work
strength. The interaction of the tip vortex and to a showed that the main features of the pressure loading of
lesser extent the inboard sheet with the fuselage thus the airframe under the direct impingement of the wake
leads to higher induced airloads and vibration levels. It and in particular the tip vortex are a large positive
was also found that the presence of the fuselage distorts pressure load due to blade passage and an equally large
the wake which in turn aects the tip-vortex strength suction peak directly under the tip vortex as indicated
and alters the nature of BVIs [186]. In addition, these above. The vortex trajectory and the pressure were
interactions can trigger a torsional aeroelastic response correlated and it was found that the suction peak
by the rotor that may lead to a premature retreating amplitude grew as the tip vortex approached, reached a
blade stall. This early work identied important new maximum, and then dropped rapidly as the vortex
phenomena which had to be accounted for in advanced collided with the surface. The pressure distribution on
models of the rotor wake. the top and the sides of the airframe was presented. The
Even in the mid-1980s, there were no computer codes accuracy of these measurements was about 10 mm; or
capable of predicting even the rst-order features of the one core diameter. The time history of the pressure
time-averaged surface pressure distribution over the during a cycle has also been investigated [193]; they note
upper surface of the simplest fuselage geometry in hover that these unsteady pressure peaks are very large and
or low-speed forward ight on the model scale. No can swamp the steady mean pressure. The presence of
models for calculating the time-dependent motion of the the fuselage aects the path of the tip vortex signicantly
468 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

blade passage eect is dominant up to about c0 301


with the inuence of the tip vortex being dominant for
the next B301: Computational results [119,195] suggest
that two physical eects are responsible for the
reduction of the suction peak. First, the presence of
axial ow within the vortex is necessary for the
reduction of the suction peak; the axial velocity
stagnates on the airframe on the advancing side and is
sucked away from the airframe on the retreating side.
The stagnation of the axial velocity forces reduction of
the suction peak. The proper time scale of the reduction,
however is obtained by allowing the deformation of the
vortex core. A simplied model of the vortex core as a
rectangular box whose aspect ratio is allowed to vary
has been developed [195] and a typical result is depicted
in Fig. 50. Here the computational results are compared
with the experimental data of [196]. Note the reduction
of the suction peak which occurs in milliseconds at the
parameters studied.
Additional experimental results [196] suggest that the
suction peak in the pressure is convected around the
retreating side of the airframe for a long period of time
after it has disappeared from the top of the airframe.
This phenomenon has been explained for a model
problem [166] and a sketch of the pressure distribution
is also shown. The computations are inviscid. On the
side of the plate where the velocity normal to the plate
and along the axis of the vortex is toward the plate, the
axial velocity stagnates on the plate leading a bulging
vortex core and a reduced suction peak as described
above. On the other hand, when the direction of the
axial velocity in the vortex is away from the plate,
Fig. 49. Comparison between computation and experiment of the vortex core thins as uid gets sucked in from the
the vortex trajectory and pressure distribution on the top of a surrounding area and the suction peak grows (Fig. 51).
model airframe for advance ratio m 0:1: The z-axis is in the Eventually, the suction peak amplitude is limited by the
direction of forward ight: (a) Tip-vortex trajectory c action of viscous forces which limit the shrinkage of
01; 301; 601; A denotes advancing side and R denotes retreating the vortex core. However, the dominant physics of the
side. (b) Pressure at c 421 on the left and c 481 on the collision process is inviscid.
right. From [195]. This eect is seen in plots of the pressure contours
around the airframe as shown in Fig. 52 at a vortex age
of 2701 [196]. Note that on the retreating side (fo0; f
refers to the angle measured from the top of the airframe
[148,194,195]. Typical results for the vortex path are and not the zero-lift angle of attack as in Fig. 3) the
shown in Fig. 49 [195]. suction peak is still strong while on the advancing side,
Based on this past work and a recent computational the pressure on the airframe at a point coinciding with
eort [195] reasonable picture of the dominant features the vortex core is actually positive. These large-scale
of the unsteady pressure on the airframe has emerged suction peak eects on the retreating side could be
under direct wake impingement conditions. For a given important in maneuvering ight and could be a factor in
rotorbody separation distance, on the top of the the fatigue life of the airframe.
airframe the blade passage eect generates a large Steady ow NavierStokes simulations of the full
positive pressure rise at c 0 (and c 2p=N for an N- wakeairframe interaction in forward ight have begun
bladed rotor), followed by a very sharp suction peak of [197]. The rotor is treated as a source term in the
the same order if the advance ratio is low enough so that governing equations; the magnitude of the sources is
the tip vortex collides with the airframe. For a vortex determined from the blade geometry and aerodynamic
whose age is c 180 c0 (i.e. for a two-bladed rotor) characteristics. The detailed boundary layer behavior
at the given rotorbody separation distance [192], the under the tip vortex as it approaches the airframe has
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 469

Fig. 50. Pressure on the airframe surface directly above the airframe centerline as the vortex approaches [195]: (a) vortex age c 2101;
(b) c 2221; (c) c 2281; and (d) c 2341: The circles indicate experimental data from [196].

also been discussed [199,200]. It is shown that ow 7. Summary


reversal occurs in the form of an eddy which grows in
time. Eventually, the boundary layer uid will wrap A comprehensive understanding of the helicopter
around the vortex uid and this may alter its eective rotor wake is elusive because the ow eld generated
strength. by a helicopter rotor is extremely complicated and
The sophistication of large-scale computational tech- dicult to measure, model and predict; experiments are
niques has improved to the point that entire helicopter expensive and also dicult to conduct. The wake and
congurations may be incorporated; results for the blade ows contain a variety of complicated and
RAH-66 Commanche have been published [198,201]. In inherently non-linear ow phenomena including dy-
Fig. 53 is depicted the streakline patterns for the V-22 namic stall, BVI, and shock interactions. Fluidstruc-
Tiltrotor operating in the helicopter mode [202]. The ture interaction problems occur throughout the rotor
V-22 is a dual purpose machine that can operate in both cycle because the rotor blades undergo pitching, apping
xed-wing and rotary-wing modes. The computation and aeroelastic motions dynamically in all regimes of
uses an overset grid approach in which a family of grid operation. Because of these complexities current models
systems each designed for accuracy in a particular region and experiments are carefully targeted in scope and
of the ow are employed. Data are communicated signicant simplications are made in order to focus on
between grids by interpolation. a specic aspect of the ow.
470 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

Fig. 52. (a) Measured pressure distribution around a model


airframe showing the dierences exhibited on the advancing
and retreating sides of the airframe [196]. Arrows denote region
below the vortex core. The angle f measures distance from the
top of the cylinder and f > 0 denotes the advancing blade side.
Xb =R measures distance along the airframe in the forward ight
direction.

Fig. 51. A sketch of the tip-vortex structure along the sides of


the airframe [166,196]. The pressure under the vortex along the
sides of the airframe is also shown. The core radius of the
vortex is greatly enlarged for clarity. Arrows denote direction of
axial ow within the vortex.

Indeed, the method employed in the calculation of the


rotor wake almost always depends on the objective of
the calculation. Helicopter design is truly an interdisci-
plinary activity and aeroelasticity of the rotor blades is
also an important feature in the modeling of the rotor
wake. On the blade, the ow can be locally supersonic
and shocks can form while downstream of the blade in
the wake, the ow is incompressible and vortical in Fig. 53. Streaklines for the V-22 Tiltrotor operating in the
character. Thus, in developing a model of the wake, helicopter mode [202].
because there are a wide variety of ow phenomena
occurring simultaneously, there is always a tradeo
between the accuracy of the model and the ease of Despite the seemingly overwhelming complexities of
incorporating the results into a comprehensive rotor the rotor wake, the past 10 years have seen great
design code which will normally include the specication advances in the understanding of isolated rotor wakes.
of blade motions for balancing moments and the elastic With the increased storage capacity and processing
properties of the blades. speed of modern computers in recent years, and the
In addition, wake computations are almost always development of sophisticated experimental techniques
accompanied by comparison with experimental data designed to measure the fully three-dimensional char-
which have been the foundation of helicopter aero- acter of rotor wakes, reasonable comparisons with
mechanics. Experimental data are used to validate a experiment may now be obtained with computational
wake model and this reects a truly coupled approach studies at low to moderate Mach number for an isolated
within the helicopter aeromechanics community: insight rotor in hover at model scale provided aeroelastic
gained from experiments is used to build a better deformations are negligible. The agreement between
theoretical and computational model for the rotor wake. computations and experiment for an isolated rotor in
Often the knowledge ow is reversed in that physical forward ight is not as good but advances in this regime
phenomena uncovered in calculations suggest appro- are also being made. In addition, drag and moment are
priate experiments. dicult to predict especially under stall conditions and
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 471

there is little experimental data to guide computational experimental section and Professor Lakshmi Sankar
approaches. provided some insights into the grid-based computa-
On the experimental side, detailed three-dimensional tions. Professors Dewey Hodges and David Peters
and unsteady measurements of the rotor wake ow are provided material for the section on aeroelasticity and
now beginning to be taken on a routine basis. However, trim. Special thanks to Henry R. Velko, a former
the amount of data to be acquired and processed in this Editor of the Journal of the American Helicopter Society
activity is staggering and eorts must be taken to who began the tradition of studying helicopter rotor
process the data in a more ecient manner. This process ows at Ohio State long before me.
is aided by the availability of post-processing tools,
normally designed for computational studies, which are
now being used to reduce experimental data. References
The long-range goal of any modeling eort is to be
able to compute the entire unsteady ow eld around a [1] Caradonna FX. The application of CFD to rotary wing
helicopter operating in any ight regime. The objective is aircraft. NASA TM 102803, 1992.
to reduce the amount of expensive full-scale testing [2] Conlisk AT. Modern helicopter aerodynamics. Annu Rev
which is now required. However, this goal will likely not Fluid Mech 1997;29:51567.
be realized for some time. This is because rotor loads [3] Gessow A. Understanding and predicting helicopter
even at model scale in the forward ight and descent behaviorFthen and now. J Am Helicopter Soc
regimes are not accurate. An additional limiting 1986;24:128.
technological barrier is the issue of noise which is very [4] Johnson W. Recent developments in rotary-wing aero-
dynamic theory. AIAA J 1986;24(8):121944.
acute at the higher ight speeds typical of modern
[5] McCroskey WJ. Vortex wakes of rotorcraft. AIAA Paper
helicopters. Noise is identied as the major impediment 1995; (95-0530).
to increased commercial use of rotorcraft. In order to [6] Gray RB. Vortex modeling for rotor aerodynamicsFthe
design quieter helicopters, accurate computation of BVI 1991 Alexander A. Nikolsky Lecture. J Am Helicopter
and high-speed impulsive noise must be made; inevitably Soc 1991;36(1):314.
this requires a more accurate computation of the ow [7] Reichert G. Helicopter aeromechanicsFintroduction
eld itself. and historical review. AGARD-LS-139, 1985. Helicopter
The extrapolation of model-scale results to the full Aeromechanics, April 1985. p. 1-11-23.
scale is often troublesome: considerable dierences [8] Phillipe JJ, Roesch P, Dequin AM, Cler A. A survey of
between full-scale ight tests and full-scale wind tunnel recent development in helicopter aerodynamics.
AGARD-LS-139, 1985. Helicopter Aeromechanics, April
tests are common [203]. Much of the disagreement
1985. p. 2-12-40.
between full-scale and model-scale experiments may be [9] Carr LW. Progress in analysis and prediction of dynamic
the lack of dynamic similarity (e.g. Reynolds numbers stall. J Aircraft 1988;25(1):617.
not matching, aeroelastic dierences, etc.). Indeed, at [10] Carr LW, Chandrasekara MS. Compressibility eects on
the current time, aeroelastic eects which are signicant dynamic stall. Prog Aerosp Sci 1996;32(6):52373.
at full scale, are most often neglected in the model-scale [11] McCroskey WJ. The phenomenon of dynamic stall.
computations. NASA TM 81264, March 1981.
Despite these challenges, with the advances in both [12] Lowson MV. Helicopter noise: analysis-prediction and
experimental and computational technology, the range methods of reduction. AGARD LS 63, 1973.
of conditions over which computations and experiments [13] Ffowcs-Williams JE, Hawkings DL. Sound generated by
turbulence and surfaces in arbitrary motion. Philos Trans
can be made will likely increase substantially in the
Roy Soc A 1969;264:32142.
future. [14] Ffowcs-Williams JE. Sound sources in aerodyna-
micsFfact and ction. AIAA J 1982;20(3):30715.
[15] Gessow A, Myers G. Aerodynamics of the helicopter.
Acknowledgements New York: Frederick Ungar, Publishing Co, 1952.
[16] McCormick BW. Aerodynamics of V/STOL ight.
The author is grateful to Dr. Thomas L. Doligalski New York: Academic Press, 1967.
who read the manuscript and provided a number of [17] Bramwell ARS. Helicopter dynamics. London: Edward
insights into the history and current state of rotorcraft Arnold, 1976. p. 408.
research. T. Alan Egolf and Dr. Brian Wake provided a [18] Stepniewski WZ, Keys CN. Rotary-wing aerodynamics,
vols. 1 and 2. New York: Dover Publications Inc, 1984.
number of insights from the industry perspective and
[19] Seddon J. Basic helicopter aerodynamics. BSP Profes-
suggested a number of changes. Hui Li suggested a sional Books, Oxford, 1990.
number of changes in an earlier version of this manu- [20] Johnson W. Helicopter theory. Princeton, NJ: Princeton
script. Santosh Kini and Vish Godavarty also provided University Press, 1980.
editing services. Professor Narayanan Komerath [21] Leishman JG. Principles of helicopter aerodynamics.
provided the basis for the material discussed in the Cambridge: Cambridge University Press, 2000.
472 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

[22] Ramachandran K, Tung C, Caradonna FX. Rotor hover conditions. Proceedings of the 44th Annual Forum of the
performance prediction using a free wake, computational American Helicopter Society, Washington, DC, 1618
uid dynamics method. J Aircraft 1989;26:110510. June 1988. p. 81932.
[23] Rankine WJ. On the mechanical principles of the action [45] Bliss DB, Teske ME, Quackenbush TR. A New
of ship propellers. Trans Inst Naval Arch 1965;6:1339. methodology for free wake analysis using curved vortex
[24] Froude RE. On the elementary relation between pitch, elements. NASA CR 3958, 1987.
slip and propulsion eciency. Trans Inst Naval Arch [46] Leishman JG, Baker A, Coyne A. Measurements of rotor
1878;19:4765. tip vortices using three-component laser doppler veloci-
[25] Glauert H. An aerodynamic theory of the airscrew. metry. J Am Helicopter Soc 1996;41(4):34253.
British ARC R&M 786, 1922. [47] Scully MP. A method of computing helicopter rotor wake
[26] Betz A. Die wichtigsten grundlagen fur den entwurf von distortion. ASRL TR 138-1, Massachusetts Institute of
luftschrauben. Z Math 1915;6:97103. Technology, June 1967.
[27] Glauert H. Airplane propellers. In: Durand WF, editor. [48] Vatistas GH, Kozel V, Mih WC. A simpler model for
Aerodynamic theory, vol. IV. Berlin: Julius Springer, concentrated vortices. Exp Fluids 1991;11:736.
1943. p. 169360. [49] Levy H, Forsdyke AG. The steady motion and stability
[28] Lock CNH. Experiments to verify the independence of a helical vortex. Proc Roy Soc A 1928;120:67090.
of the elements of an airscrew blade. British R&M 853, [50] Widnall SE. The stability of a helical vortex lament.
1924. J Fluid Mech 1972;54(4):64163.
[29] Fogarty LE, Sears WR. Potential ow around rotating [51] Gupta BP, Loewy RG. Theoretical analysis of the
advancing cylindrical blade. J Aerosp Sci 1950;17(9):599. aerodynamic stability of multiple, interdigitated helical
[30] Loewy RG. A two-dimensional approximation to the vortices. AIAA J 1974;12(10):13817.
unsteady aerodynamics of rotary wings. J Aerosp Sci [52] Landgrebe AJ. The wake geometry of a hovering rotor
1957;24(2):8192. and its inuence on rotor performance. J Am Helicopter
[31] Eppler R. Airfoil design and data. Berlin: Springer, 1990. Soc 1972;17(4):315.
[32] Katz J, Plotkin A. Low-speed aerodynamics: from wing [53] Kocurek J, Tangler JL. A prescribed wake lifting surface
theory to panel methods. New York: Mcgraw-Hill, 1991. hover performance analysis. J Am Helicopter Soc
[33] Bertin JJ, Smith ML. Aerodynamics for engineers, 2nd 1977;22(1):2435.
ed. Englewood Clis: Prentice-Hall, 1989. [54] Bliss DB, Wachspress DA, Quackenbush TR. A new
[34] Glauert H. Elements of airfoil and airscrew theory, 2nd approach to the free wake problem for hovering rotors.
ed. London: Cambridge University Press, 1948. Proceedings of the 41st Annual Forum of the American
[35] Goldstein S. On the vortex theory of screw propellers. Helicopter Society, May 1985.
Proc Roy Soc London A 1929;112:44065. [55] Clark DR, Leiper AC. The free wake analysis: a method
[36] Lock CNH. The application of Goldsteins airscrew for the prediction of helicopter rotor hovering perfor-
theory to design. British R&M 1377, 1930. mance. Presented at the 25th Forum of the American
[37] Jenney DS, Olson JR, Landgrebe AJ. A reassessment of Helicopter Society, 1969.
rotor hovering performance prediction methods. 23rd [56] Sadler GS. Development and application of a method for
Forum of the American Helicopter Society, May 1969. predicting rotor free wake position and resulting rotor
[38] Gray RB. On the motion of the helical vortex shed from a blade loads. NASA CR-1911, vols. I and II, December
single-bladed hovering helicopter rotor and its applica- 1971.
tion to the calculation of the spanwise aerodynamic [57] Miller WO, Bliss DB. Direct periodic solutions of rotor
loading. Princeton University Aero Engr Dept, Report free wake calculations. J Am Helicopter Soc
No. 313, September 1955. 1993;38(2):5360.
[39] Gray RB. An aerodynamic analysis of a single-bladed [58] Quackenbush TR, Bliss DB, Wachpress DA. Free wake
rotor in hovering and low-speed forward ight as analysis of hover performance using a new inuence
determined from smoke studies of the vorticity distribu- coecient method. NASA CR-4309, 1990.
tion in the wake. Princeton University Aero Engr Dept, [59] Bagai A, Leishman JG. Rotor free-wake modeling using a
Report No. 356, September 1956. pseudo-implicit technique-including comparisons with
[40] Batchelor GK. Introduction to uid dynamics. Cam- experimental data. J Am Helicopter Soc 1995;40:2941.
bridge: Cambridge University Press, 1967. [60] Tinoco EN, Johnson FT, Freeman LM. Application of a
[41] Ting L, Tung C. The motion and decay of a vortex in a higher order panel method to realistic supersonic cong-
uniform stream. Phys Fluids 1965;8:103951. urations. J Aircraft 1980;17(4):3844.
[42] Moore DW, Saman PG. The motion of a vortex [61] Hess JL. Panel methods in computational uid dynamics.
lament with axial ow. Philos Trans Roy Soc Ann Rev Fluid Dyn 1990;22:25574.
1972;272:40329. [62] Morino L, Gennaretti M. Boundary integral equation
[43] Widnall SE, Bliss DB, Zalay A. Theoretical and experi- methods for aerodynamics. In: Atluri SN, editor. Com-
mental study of the stability of a vortex pair. In: Olsen J, putational non-linear mechanics in aerospace engineering,
Goldberg A, Rogers M, editors. Aircraft wake turbulence vol. 147, Progress in Astronautics and Aeronautics,
and its detection. New York: Plenum Press, 1971. American Institute of Aeronautics and Astronautics,
p. 30538. Washington, DC, 1991.
[44] Egolf TA. Helicopter free-wake prediction of complex [63] Van Dyke M. Perturbation methods in uid mechanics,
wake structures under bladevortex interaction operating anotated edition. Stanford: Parabolic Press, 1975.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 473

[64] Leishman JG, Beddoes TS. A semi-empirical model [84] Kini S, Conlisk AT. The development of steady state
for dynamic stall. J Am Helicopter Soc 1989;34(3): rotor wakes. American Helicopter Society Aeromecha-
317. nics Specialists Meeting, Atlanta, GA, 1314 November
[65] Betz A. Behaviour of vortex systems. NACA-TM-713, 2000.
1932. [85] Caradonna F, Hendley E, Silva M, Huang S, Komerath
[66] Johnson W. A lifting-surface solution for vortex-induced N, Reddy U, Mahalingam R, Funk R, Ames R, Darden
airloads. AIAA J 1971;9(4):68995. L, Villareal L, Gregory, Wong O. Performance char-
[67] Schlichting H, Thomas HHBM. Note on the calculation acteristics of a model rotor in axial ight. J Am
of the lift distribution of swept wings. Royal Aircraft Helicopter Soc 1999;44(2):1018.
Establishment Report No. Aero. 2236, 1947. [86] Tannehill JC, Anderson DA, Pletcher RH. Computa-
[68] Zachmanoglou EC, Thoe DW. Introduction to partial tional uid mechanics and heat transfer, 2nd ed.
dierential equations with applications. New York: Washington: Taylor & Francis, 1997.
Dover, 1986. [87] Ahmad J, Duque EPN. Helicopter rotor blade computa-
[69] Morino L, Chen LT, Suciu EO. Steady and oscillatory tion in unsteady ows using moving embedded grids.
subsonic and supersonic aerodynamics around complex J Aircraft 1994;33(1):5460.
congurations. AIAA J 1975;13(3):36874. [88] Bangalore A, Sankar LN. Forward ight analysis of
[70] Lee YJ, Yang JY. A panel method for arbitrary moving slatted rotors using NavierStokes methods. AIAA Paper
boundary problems. AIAA J 1990;28(3):4328. 96-0675, 1996.
[71] Maskew B. Prediction of subsonic aerodynamic char- [89] Biswas R, Strawn RC. A new procedure for dynamic
acteristics: a case for low-order panel methods. J Aircraft adaptation of three-dimensional unstructured grids. App
1982;19(2):15763. Numer Math 1994;13:43752.
[72] Carlin GJ, Staedell WE, Hodges RM. Analysis of V-22 [90] Wake BE, Sankar NL. Solutions of the NavierStokes
tilt-rotor aircraft using panel methods. Proceedings of the equations for the ow about a rotor blade. J Am
42nd Annual Forum of the American Helicopter Society, Helicopter Soc 1989;34(2):1323.
Washington, DC, June 1986. p. 30314. [91] Srinivasan GR, Baeder JD, Obayashi S, McCroskey WJ.
[73] Katz J, Maskew B. Unsteady low-speed aerodynamic Floweld of a lifting rotor in hover: a NavierStokes
model for complete aircraft congurations. J Aircraft simulation. AIAA J 1992;30(10):23718.
1987;25(4):30210. [92] Beam RM, Warming RF. An implicit factored scheme
[74] Saman PG. Vortex dynamics. London: Cambridge for the compressible NavierStokes equations. AIAA J
University Press, 1992. 1978;(16):393401.
[75] Phillips WF, Snyder DO. Modern adaptation of Prandtls [93] Jameson A, Yoon S. Lowerupper implicit schemes with
lifting-line theory. J Aircraft 2000;37(4):66270. multiple grids for the Euler equations. AIAA J
[76] Hui L, Burggraf OR, Conlisk AT. On the formation of 1987;25(7):35772.
a rotor Tip-vortex. American Helicopter Society Aero- [94] Yoon S, Jameson A. An LU-SSOR scheme for the Euler
mechanics Specialists Meeting, Atlanta, GA, 1314 and NavierStokes equations. AIAA Paper 87-0600,
November 2000. January 1987.
[77] Peters DA, Boyd DD, Cheng JH. Finite-state induced- [95] Yee HC. A class of high-resolution explicit and implicit
ow model for rotors in hover and forward ight. J Am shock-capturing methods. Computational Fluid Dy-
Helicopter Soc 1989;34(4):517. namics, Rhode-St. Genese, Belgium, 610 March 1989.
[78] Peters DA, Cheng JH. Correlation of measured induced [96] Steger JL, Warming RF. Flux vector splitting of the gas
velocities with a nite-state wake model. J Am Helicopter dynamic equations with application to nite dierence
Soc 1991;36(3):5970. methods. J Comput Phys 1981;40:26393.
[79] Su A, Yoo KM, Peters DA. Extension and validation of [97] van Leer B. Flux vector splitting for the Euler equations.
an unsteady wake model for rotors. J Aircraft Proceedings of the Eighth International Conference on
1992;29(3):37483. Numerical Methods in Fluid Dynamics, Lecture Notes in
[80] Egolf TA, Landgrebe AJ. Helicopter rotor wake geome- Physics, vol. 170. Berlin: Springer, 1982. p. 50712.
try and its inuence in forward ight, vol. 1: generalized [98] Roe PL. Approximate Rieman solvers, parameter vec-
wake geometry and wake eect on rotor airloads and tors, and dierence schemes. J Comput Phys 1981;
performance. NASA CR 3726, October 1983. 43:3549.
[81] Leishman JG, Bagai A. Challenges in understanding the [99] Kreyszig E. Advanced engineering mathematics, 7th ed.
vortex dynamics of helicopter rotor wakes. AIAA J New York: Wiley, 1993.
1998;36(7):113040. [100] van Leer B. Toward the ultimate upwind dierence
[82] Torok MS, Berezin CR. Aerodynamic and wake meth- scheme, II: monotonicity and conservation combined in a
odology evaluation using model UH-60A experimental second Order scheme. J Comput Phys 1972;14:36170.
data. J Am Helicopter Soc 1993;39(2):219. [101] Harten A. High resolution schemes for hyperbolic
[83] Jain R, Conlisk AT. Interaction of tip-vortices in the conservation laws. J Comput Phys 1983;49:35793.
wake of a two-bladed rotor in axial Flight. J Am [102] Harten A, Engquist B, Osher S, Chakravarthy S.
Helicopter Soc 2000;45(3):15764 [an earlier version of Uniformly high order accurate non-oscillatory schemes.
this paper was presented at the 53rd Forum of the J Comput Phys 1987;71:231303.
American Helicopter Society, Washington, DC, 2022 [103] Hariharan N, Sankar LN. Higher order numerical
May 1998]. simulation of rotor ow eld. 42nd Annual Forum of
474 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

the American Helicopter Society, Washington, DC, 1113 [120] Sankar LN, Wake BE, Lekoudis SG. Solution of the
May 1994. unsteady Euler equations for xed and rotor wing
[104] Harten A. On a class of high resolution total-variation- congurations. J Aircraft 1986;23(4):2839.
stable nite dierence schemes. SIAM J Numer Anal [121] Agarwal RK, Deese JE. Euler calculations of a helicopter
1979;54:10136. rotor in hover. J Aircraft 1987;24(4):2318.
[105] Wake BE, Choi D. Investigation of high-order upwinded [122] Liu CH, Thomas JL, Tung C. NavierStokes calculations
dierencing for vortex convection. AIAA J for the vortex wake of a rotor in hover. AIAA Paper
1996;34(2):3327. 83-1676, July 1983.
[106] Wake BE, Egolf TA, Choi D. Resolution and convection [123] Chen CS, Velko HR, Tung C. Free-wake analysis of a
of tip-vortex using higher-order methods. Int J Comput rotor in hover. AIAA Paper 87-1245, AIAA Fluid
Fluid Dyn, 2001, in review. Dynamics, Plasma Dynamics and Lasers Conference,
[107] Caradonna FX, Phillipe JJ. The ow over a helicopter Honolulu, 810 June 1987.
blade tip in the transonic regime. Vertica 1978; [124] Srinivasan GR, McCroskey WJ. NavierStokes calcula-
2:4360. tions of hovering rotor owelds. J Aircraft 1988;25:
[108] Strawn RC, Caradonna FX. Conservative full potential 86574.
method for unsteady transonic rotor ows. AIAA J [125] Srinivasan GR, Baeder JD. Recent advances in Euler and
1987;25:1938. NavierStokes methods for calculating helicopter rotor
[109] Sankar LN, Prichard D. Solution of transonic ow past aerodynamics and acoustics. Fourth International Sym-
rotor blades using the conservative full potential equa- posium on Computational Fluid Dynamics, Davis, CA,
tion. AIAA Paper 85-5012, October 1985. 912 September 1991.
[110] Egolf TA, Sparks SP. A full potential rotor analysis with [126] Wake BE, Baeder JD. Evaluation of a NavierStokes
wake inuence using an innerouter domain technique. analysis method for hover performance prediction. J Am
Proceedings of the 42nd Annual Forum of the American Helicopter Soc 1996;41(7):717.
Helicopter Society, 1986. p. 9971011. [127] Srinivasan GR, Baeder JD. TURNS: a free-wake Euler
[111] Chang IC, Tung C. Numerical solution of the full- NavierStokes numerical method for helicopter rotors.
potential equation for rotors and oblique wings using a AIAA J 1993;31(5):95962.
new wake model. AIAA Paper 85-0268, 1985. [128] Rajagopalan RG, Mathur SR. Three dimensional analy-
[112] Arieli R, Tauber ME. Computation of subsonic and sis of a rotor in forward ight. J Am Helicopter Soc
transonic ow about lifting rotor blades. AIAA Paper 79- 1993;38(3):1425.
1667, August 1979. [129] Steinho J. Vorticity connement: a new technique for
[113] Caradonna FX, Tung C. Experimental and analytical computing vortex dominated ows. In: Caughey DA,
studies of a model helicopter rotor in hover. NASA TM Hafez MM, editors. Frontiers of computational uid
81232, September 1981. dynamics. Chichester: Wiley, 1994.
[114] Ramachandran K, Schlechtriem S, Caradonna FX, [130] Wenren Y, Steinho J. Application of vorticity conne-
Steinho J. The application of vorticity embedding to ment to the prediction of the wake of helicopter rotors
the computation of advancing rotor ows. 49th Annual and complex bodies. AIAA 99-3200, 17th Applied
Forum of the American Helicopter Society, 1921 May Aerodynamics Conference, Norfolk, VA, June 28July 1,
1993. 1999.
[115] Landgrebe AJ. New directions in rotorcraft computa- [131] Tang L, Baeder JD. Improved Euler simulations of
tional aerodynamics research. US 75th AGARD Fluid hover tip vortices with validation. Proceedings of
Dynamics Panel Meeting on Aerodynamics and Aero- the 55th Annual American Helicopter Society Forum,
acoustics of Rotorcraft, Berlin, October 1994. 1999.
[116] Srinivasan GR, Sankar LN. Status of Euler and Navier [132] Hodges DH. Review of composite rotor blade modeling.
Stokes CFD methods for helicopter applications. Pre- AIAA J 1990;28:5615.
sented at the American Helicopter Society Specialists [133] Bauchau OA. A beam theory for anisotropic materials.
Meeting on Aeromechanics Technology and Product J Appl Mech 1985;52:41622.
Design, Bridgeport, CT, 1113 October 1995. [134] Padeld GD. Helicopter ight dynamics: the theory and
[117] Renzoni P, DAlascio A, Kroll N, Peshkin D, Hounjet application of ying qualities and simulation modeling.
MHL, Boniface J-C, Vigevano L, Allen CB, Badcock K, AIAA Education Series, 1996.
Mottura L, Scholl E, Kokkalis A. EROSFa common [135] Bielawa RL. Rotary wing structural dynamics and
European Euler code for the analysis of the helicopter aeroelasticity. AIAA Education Series, 1992.
rotor oweld. Prog Aerosp Sci 2000;36(56): [136] Dowell EH, Curtiss Jr HC, Scanlan RH, Sisto F, editors.
43785. A modern course in aeroelasticity. Boston: Kluwer, 1989.
[118] Sankar LN, Tung C. Euler calculations for rotor [137] Loewy RG. Review of rotary-wing, V/STOL dynamic
congurations in unsteady forward ght. 42nd Annual and aeroelastic problems. J Am Helicopter Soc
Forum of the American Helicopter Society, Washington, 1969;14:323.
DC, 24 June 1986. [138] Peters DA. Flap-lag stability of helicopter rotor blades in
[119] Tung C, Lee S. Evaluation of hover prediction codes. forward ight. J Am Helicopter Soc 1975;20:213.
Proceedings of the 50th Annual Forum of the American [139] Wei F-S, Peters DA. Lag damping in autorotation by a
Helicopter Society, Washington, DC, 1113 May 1994. perturbation method. 34th Forum of the American
p. 82944. Helicopter Society, May 1978.
A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476 475

[140] Ormiston RA, Hodges DH. Linear ap-lag dynamics of Annual Forum of the American Helicopter Society,
hingeless helicopter rotor blades in hover. J Am 2022 May 1998.
Helicopter Soc 1972;17:214. [159] Funk R, Fawcett PA, Komerath NM. Instantaneous
[141] Ganguli R, Chopra I. Aeroelastic optimization of an velocity elds in a rotor wake by spatial correlation
advanced geometry helicopter rotor. J Am Helicopter Soc velocimetry. AIAA Paper 93-3081, 24th Fluid Dynamics
1996;41:1828. Conference, Orlando, FL, July 1993.
[142] Friedmann FP, Hodges DH. Rotary-wing aeroelasticity [160] Heineck JT, Yamauchi GK, Wadcock AJ, Lourenco L,
with application to VTOL vehicles. In: Noor AK, Abrego AJ. Application of three-component PIV to a
Venneri SL, editors. Flight-vehicle materials, structures, hovering rotor wake. Presented at the 56th Annual
and dynamics, vol. 5. New York: ASME, 1992. p. 299 Forum of the American Helicopter Society, May 2000.
391. [161] Gorton S, Meyers J, Berry JD. Velocity measurements
[143] Kunz DH. Survey and comparison of engineering beam near the empennage of a small-scale helicopter model.
theories for helicopter rotor blades. J Aircraft Presented at the 51st Annual Forum of the American
1994;31:4739. Helicopter Society, May 1995.
[144] Lorber PF. Aerodynamic results of a pressure-instrumen- [162] Chen RTN. A survey of non-uniform inow models for
ted model rotor test at the DNW. Journal of American rotorcraft ight dynamics and control applications.
Helicopter Society 1991;36(4):6676. Vertica 1990;14:14790.
[145] Parthasarathy SP, Cho YI, Black LH. Wide-eld [163] Elliott JW, Altho SL. Inow measurement made with a
shadowgraph ow visualization of tip vortices generated laser velocimeter on a helicopter model in forward ight.
by a helicopter rotor. AIAA 85-1557, July 1985. NASA TM 100541-100545, vols. IV, April 1988.
[146] Norman TR, Light JS. Rotor tip vortex geometry [164] Hoad DR, Altho SL, Elliott JW. Rotor inow
measurements using the wide-eld shadowgraph techni- variability with advance ratio. Proceedings of the 44th
que. J Am Helicopter Soc 1987;32(2):4050. Annual Forum of the American Helicopter Society,
[147] Felker FF, Light JS. Aerodynamic interactions between a Washington, DC, 1988. p. 5772.
rotor and wing in hover. J Am Helicopter Soc [165] Hoad DR. Rotor induced-inow-ratio measurements
1988;33(2):5361. and CAMRAD calculations. NASA TP 2946, January
[148] Bagai A, Leishman JG. Improved wide-eld shadow- 1990.
graph set-up for rotor wake visualization. J Am [166] Lee JA, Burggraf OR, Conlisk AT. On the impulsive
Helicopter Soc 1992;37(3):8692. blocking of a vortex-jet. J Fluid Mech 1998;369:
[149] Landgrebe AJ, Johnson BV. Measurements of model 30131.
helicopter rotor ow velocities with a laser doppler [167] Panaras AG. Numerical modeling of the vortex/airfoil
velocimeter. J Am Helicopter Soc 1974;19(3):3943. interaction. AIAA J 1987;25(1):511.
[150] Desopper A, Lafon P, Ceroni P, Phillippe JJ. Ten years of [168] Poling DR, Wilder MC, Telionis, DP. Fundamental
rotor ow studies at ONERA: state of the art and future research in helicopter rotor bladevortex interaction
studies. Proceedings of the 42nd Annual Forum of the modeling. Proceedings of the 41st Annual Forum of the
American Helicopter Society, Washington, DC, June American Helicopter Society, vol. 1, 1991. p. 42133.
1986. p. 26777. [169] Srinivasan GR, McCroskey WJ. Numerical simulations
[151] Liou SG, Komerath NM, Lal MK. Measurement around of unsteady airfoilvortex interactions. Vertica 1987;
a rotor blade excited in pitch, Part 1: dynamic inow. 11(1/2):328.
J Am Helicopter Soc 1994;39(2):312. [170] Yu Y. Rotor bladevortex interaction noise. Prog Aerosp
[152] Kittleson JK, Yu Y. Transonic rotor ow measurement Sci 2000;36(2):97115.
technique using holographic interferometry. J Am Heli- [171] Yu YH. Rotor bladevortex interaction noise: generating
copter Soc 1985;30(4):310. mechanisms and its control concepts. American Helicop-
[153] Thompson TL, Komerath NM, Gray RB. Visualization ter Society Specialists Meeting on Aeromechanics
and measurement of the tip vortex core of a rotor blade in Technology and Product Design, Bridgeport, CT, 1113
hover. J Aircraft 1988;25(12):111321. October 1995.
[154] Mba MN, Meylan C, Mansca C, Favier D. Radial [172] Hassan AA, Charles BD, Tadghighi H, Burley CL. A
distribution of circulation of a rotor in hover measured by consistent approach for modelling the aerodynamics of
a laser velocimeter. Proceedings of the 10th European self-generated rotor bladevortex interactions. American
rotorcraft Forum, The Hague, Netherlands, August 1984. Helicopter Society 49th Annual Forum, 1921 May
[155] Kim JM, Komerath NM. Summary of the interaction 1993.
of a rotor wake with a circular cylinder. AIAA J [173] Gallman JM, Tung C, Schultz KJ, Splettstoesser W,
1995;33:4708. Buchholz H, Speigel P, Burley CL, Brooks TF, Boyd DD.
[156] Ghee T, Berry JD, Zori LAJ, Elliott JW. Wake geometry Eect of wake structure on bladevortex interaction
measurements and analytical calculations on a small-scale phenomena: acoustic prediction and validation. Presented
rotor model. NASA TP 3584, 1995. at the First Joint CEASAIAA Aeroacoustic Conference,
[157] McAlister KW, Schuler CA, Branum L, Wu JC. 3-D Munich, June 1995.
measurements near a hovering rotor for determining [174] Heller H, Schultz KJ, Ahmed SR. Unsteady surface
prole and induced drag. NASA TP 3577, August 1995. pressure characteristics on helicopter blades: a key to the
[158] Bhagwat MJ, Leishman GJ. On the relationship between physics of rotor noise. 19th ICAS Congress, Paper 94-
blade circulation and tip vortex characteristics. 54th 2.7.2, Anaheim, CA, September 1994.
476 A.T. Conlisk / Progress in Aerospace Sciences 37 (2001) 419476

[175] Yu YH. Miss distance for rotor bladevortex interaction [190] Berry JD, Altho SL. Inow velocity perturbations due
noise reduction. AIAA Paper 96-1738, Second AIAA/ to fuselage eects in the presence of a fully interactive
CEAS Aeroacoustics Conference, State College, PA, 68 wake. Proceedings of the 46th Annual Forum of The
May 1996. American Helicopter Society, Washington, DC, May
[176] Hassan AA, Tung C, Sankar LN. Euler solutions for self- 1990. p. 111120.
generated rotor bladevortex interactions. Int J Numer [191] Liou SG, Komerath NM, McMahon HM. Velocity
Meth Fluids 1992;15:42751. measurements of airframe eects on a rotor in low-speed
[177] Burley CL, Tadghighi H. Importance of high accuracy forward ight. J Aircraft 1989;26(4):3408.
blade motion and airloads prediction for acoustic [192] Brand AG, McMahon HM, Komerath NM. Surface
analysis. American Helicopter Society 50th Annual pressure measurements on a body subject to vortex wake
Forum, 1113 May 1994. interaction. AIAA J 1989;27:56974.
[178] Brooks TF, et al. Analysis of a higher harmonic control [193] Leishman JG, Bi N. Aerodynamic interactions between a
test to reduce blade vortex interaction noise. J Aircraft rotor and a fuselage in forward ight. J Am Helicopter
1994;31(6):13419. Soc 1990;35:2231.
[179] Srinivasan GR, McCroskey WJ. Euler calculations of [194] Aes H, Conlisk AT, Kim JM, Komerath NM. Model
an advancing rotor with a line vortex. AIAA J for rotor tip vortexairframe interaction, Part 2: compar-
1993;31(9):165966. ison with experiment. AIAA J 1993;31(12):227482.
[180] Yu Y, Gmelin B, Splettstoesser W, Philippe J, Prieur J, [195] Radcli TD, Burggraf OR, Conlisk AT. On the three-
Books T. Reduction of helicopter bladevortex interac- dimensional interaction of a rotortip vortex with a
tion noise by active rotor control. Prog Aerosp Sci cylindrical surface. J Fluid Mech 2000;425:30134.
1997;33(910):64787. [196] Kim JM, Komerath NM. Summary of the interaction
[181] Sheridan PF, Smith RF. Interactional aerodynamicsFa of a rotor wake with a circular cylinder. AIAA J
new challenge to helicopter technology. J Am Helicopter 1995;33:4708.
Soc 1980;25(1):321. [197] Zori LAJ, Rajagopalan RG. NavierStokes calculations
[182] Prouty RW, Amer KB. The YAH-64 empennage and of rotorairframe interaction in forward ight. J Am
tail rotorFa technical history. Presented at the 38th Helicopter Soc 1995;40(2):5767.
Annual Forum of the American Helicopter Society, [198] Duque EPN, Dimanlig ACB. NavierStokes simulation
47 May. of the AH-66(commanche) helicopter. Presented at the
[183] Wilson JC, Mineck RE. Wind-tunnel investigation of 1994 American Helicopter Society Aeromechanics Spe-
helicopter rotor wake eects on three helicopter fuselage cialists Conference, San Francisco, CA, 1921 January
models. NASA TM X-3185, March 1975. 1994.
[184] Landgrebe AJ, Mott RC, Clark DR. Aerodynamic [199] Xiao Z, Aes H, Conlisk AT. The boundary layer ow
technology for advanced rotorcraft. J Am Helicopter Soc due to a vortex approaching a cylinder. J Fluid Mech
1977;22(2 and 3). 1994;275:3358.
[185] Smith CA, Betzina MD. Aerodynamic loads induced by a [200] Xiao Z, Burggraf OR, Conlisk AT. The interacting
rotor on a body of revolution. J Am Helicopter Soc boundary layer due to a vortex outside a cylinder. J Fluid
1986;31(1):415. Mech 1997;346:31946.
[186] Smith CA. Some eects of wake distortion due to fuselage [201] Duque EPN, Berry JD, Dimanlig ACB, Budge AM. A
ow eld on rotor thrust limits. Army Research Oce comparison of computed and experimental owelds of
Workshop on Rotor Wake Technology, Raleigh, NC, the RAH-66 Helicopter. Presented at the American
April 1979. Helicopter Society Specialists Meeting on Aeromecha-
[187] Crouse GL, Leishman JG. Interactional aerodynamic nics Technology and Product Design, Bridgeport, CT,
eects on rotor performance in hover and forward ight. 1113 October 1995.
Proceedings of the 48th Annual Forum of The American [202] Meakin RL. Moving body overset grid methods for
Helicopter Society, Washington, DC, 35 June 1992. complete tiltrotor simulations. AIAA Paper 93-3350-CP,
p. 51323. Presented at the 11th AIAA Computational Fluid
[188] Lorber PF, Egolf TA. An unsteady rotorfuselage Dynamics Conference, Orlando, FL, July 1993.
interaction analysis. J Am Helicopter Soc 1990; [203] Tung C, Bousman WG, Low S. A comparison of airload
35(7):3242. data between model-scale rotor and full-scale ight test.
[189] Quackenbush TR, Lam C-MG, Bliss DB. Vortex American Helicopter Society Specialists Meeting on
methods for the computational analysis of rotor/body Aeromechanics Technology and Product Design, Bridge-
interaction. J Am Helicopter Soc 1994;39(4):1424. port, CT, 1113 October 1995.

You might also like