You are on page 1of 24

57

Vol. IXb, 1958

Unsteady Flow Through Helicopter Rotors 1)


By

HOLT ASHLEY, EUGENE BRUNELLE, a n d H E R B E R T H . MOSER,

Cambridge, Mass., USA 2)


1. I n t r o d u c t i o n

The related problems of aerodynamically-induced vibration and aeroelastic


instability are a regular source of annoyance and danger to designers of rotary
wing aircraft, but most past attempts to cope with them have involved ruleof-thumb modifications to the rotor and its linkages. It has been found, for
example, that mass-balancing the rotor blade with respect to its quarter-chord
line, which is done to minimize control forces, normally assures that flutter
speeds lie outside of the operating range. Theoretical analyses of vibration
phenomena generally put more emphasis on accurately representing the mechanical properties of the rotor than on finding reliable expressions for the
aerodynamic loads. As a rule, some adaptation of steady-flow theory has been
employed, often on a two-dimensional basis without regard for the several
sources of aerodynamic induction present in the actual rotor flow field.
As in the case of fixed-wing aircraft, however, post-war efforts to push the
performance of the lifting rotor closer to its inherent limits are creating, among
many other requirements, a need for more rational, precise aerodynamic tools.
The ultimate obiective is clearly a complete theory of three-dimensional,
unsteady flow through the rotor in any flight condition. This objedtive is far
from achievement today, for reasons that need not be elaborated to anyone
familiar with unsteady lifting-surface theory. Nevertheless, a significant forward step was recently made by LOEWY [3] 8) (and independently by TIMMAN
and VAN DE VOOREN [5]), who was first able to explain the large loss of aerodynamic damping experienced by a hovering rotor when it vibrates at or near
frequency ratios m which are integral multiples of the number of blades.
LOEWY'S theoretical model, which is discussed further in subsequent sections,
is purely two-dimensional but accounts for the influence of the counter-vortices
that are shed from the trailing edge and spiral back into close proximity with
the blades because of the helical shape of the wake.
1) This investigation was performed under USAF Contract AF33(616)-3270 sponsored jointly
by the US Army and Dynamics Branch of the'Aircraft Laboratory, "~Vright Air Development
Center.
3) Department of Aeronautical Engineering, Massachusetts Institute of Technology.
3) Numbers in brackets refer to References, page 78.

58

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H. MOSER

ZAMI"

The success and mathematical simplicity of LoEwY's work give strong encouragement to other similar unsteady-flow investigations. The present paper
undertakes to review two such, which were carried out incidentally to a broad
program of research on helicopter vibrations at the Massachusetts Institute of
Technology. One, a two-dimensional study of ground effect on an oscillating
rotor blade, is described in section 3. The other a first and rather unwieldy
attempt to predict unsteady three-dimensional loading of a blade in simple
harmonic motion by analogy with REISSNER'S theory [12] of oscillating wings,
is covered in section 4. By way of introduction to the important influence of
unsteady effects, section 2 discusses the comparison between two-dimensional
calculations and measurements of the forced response of a simplified rotor
model.
Principal Symbols
a

b'
bo
C(k) = F + i G
C ' = F ' + iG'
CF.D.

Cp.D.
h' = 2 ~z u/Y2
n'o
H~'I
i = ]/-- 1

In, ]n, K~
k = w b'/V = m/r
t
ko
L,
m = ~o/Y2
n

P
r

r', ~f , r*, ~7"


R, R~
S

t
U
V=~r

W,#

location of pitch axis, in semichords aft of midchord axis;


blade semichord;
semichord at midspan of blade;
THEOOORSEN'S lift-deficiency function [2] ;
LOEWY'S lift-deficiency function [3];
flap-damping coefficient;
pitch-damping coefficient;
spacing between successive rows of helical wake;
distance between rotor plane and ground plane;
Hankel function of the second kind and order n;
imaginary unit;
Bessel functions, as defined by WATSON [141 ;
reduced frequency of oscillation;
reduced frequency based on b'o;
modified Struve function of order n;
frequency ratio ;
physical rotor revolution index;
functions defined by equations (22), (23), and (24);
ambient pressure, total number of shed wakes;
radius of blade section;
radial coordinates measured from blade midspan (Figure 9) ;
radius of rotor;
dimensionless semispan of blade;
time coordinate;
inflow velocity to rotor;
upwash at blade surface;
' free-stream' velocity;
wake weighting functions;

Vol. IXb, 1958

X', ~1

x,~,z,~
Y

Zm
r

F
/,,
0,~o
~,$
0)
(L~z
{0 0

.c2
a
n

( )("/
(-)

Unsteady Flow Through Helicopter Rotors

59

dimensional streamwise coordinates, positive aft;


dimensionless streamwise coordinates;
vertical displacement of blade section;
dimensionless midchord sweep (Figure 9).
angle of attack;
spanwise and chordwise components of running circulation
of vortex sheet;
total circulation around airfoil;
circulation function defined by equation (18) ;
angle variables defined by equation (32) ;
dummy variables of integration in streamwise direction;
density of incompressible fluid;
circular frequency of simple harmonic oscillation;
frequency of vertical displacement of rotor hub;
frequency of collective pitch Variation;
angular velocity of rotor.
Denotes a quantity on the blade surface;
rotor revolution index; order of function;
denotes a quantity in the shed wake;
denotes a dimensional quantity;
two-dimensional flow;
complex amplitude of a quantity with simple harmonic
time dependence.

2. Forced R e s p o n s e of a Hovering Rotor Calculated by A e r o d y n a m i c

Strip Theory
In a recently issued report [11, one of the present authors and co-workers
describe an extensive study of the dynamic response of a helicopter rotor subjected to several types of forcing. The determination and mathematical description of the airloads, both steady and unsteady, was based on the wellknown strip-theory approach, in which the flow at any given blade station is
assumed unaffected by that at any other station. The application of this assumption to the case of a hovering rotor is, of course, equivalent to describing
the flow at any blade station as two-dimensional at all azimuth locations.
The above-mentioned report considers, in particular, two cases. In the first
case, the response of the first two natural modes, rigid flapping and first
bending, of a fuUy-articulated, two-bladed rotor to a sinusoidal variation in
collective pitch was investigated. Equations of motion were derived by an
energy approach and solved for a range of frequency ratios ~o0/~Q from 0 to 4,

60

HOLT ASHLEY', EUGENE BRUNELLE~ and HERBERT H. MOSER

ZAMP

and for a rotor corresponding to that used in the experimental phase. In defining the natural modes, the rigid flapping mode shape was, of course, known
and the firstbending mode was described by the common approximation
y(~) _~ 4 ~

3 7,

where ~ is the station radius measured from the hub, y(~) is the displacement
from its equilibrium position of the blade section located by ~/, and R is the
radius of the rotor.
The second case considered was the response of the first two modes to a
sinusoidal vertical displacement of the rotor hub. These equations were derived
by GALERKIN'S method, and the solution was obtained for the frequency
ratios c%/f2 from 0 to 4, and for the same rotor as in the first case.
The primary purpose of the investigation was to measure the effects of the
unsteady aerodynamic forces on the rotor. Accordingly, the rotor response to
each type of forcing was computed by each of three existing unsteady aerodynamic theories, and the predicted response corresponding to each theory
was compared with the measured response obtained in the experimental
program.
The first unsteady aerodynamic theory used was the so-called' quasi-static'
theory in which the effect of the wake upon the aerodynamic forces acting on
the airfoil is neglected entirely. The second theory consisted of assuming that
the shed vortices arising from the varying circulation about the airfoil trail off
along the chord-line of the airfoil to infinity. This assumption gives rise to the
'lift-deficiency function' C(k) described by THEODORSEN [2].
The third theory considered, one applicable to helicopter rotors in particular, consists of assuming, and accounting for, the presence of not only the
shed vortices trailing off behind the airfoil, but also those lying in planes
(approximately) in or beneath the plane of rotation due to preceding blades
and/or revolutions. This is the previously mentioned method of LOEWY [3],
which was first verified experimentally by DAUGHDAY,DUWALT,and GATES[4].
A similar theoretical model was used by TIMMAN and VAN DE VOOREN [5],
but these authors considered only the particular case of the vortex pattern in
the plane of rotation (corresponding to no inflow through the rotor). Applying
this theory in the expressions for lift in the equations of motion consists of
simply replacing C(k) by a 'modified lift-deficiency' function C'(k, m, h')
which depends not only on the reduced frequency k, but also on the inflow
and number of blades.
To provide background for the developments that follow, some results of
the above work are reproduced here in Figures 1, 2, and 3. Figures 1 and 2
present the magnitude and phase, respectively, of the flapping response to
sinusoidal variation in collective pitch; Figure 3 shows the magnitude of the

Unsteady Flow Through Helicopter Rotors

Vol. I X b , 1958

61

1"2

1'0

0'8
o

0"6

0"~.

0"2
i

0:5

1.0

1"5 to o

2"0

2"5

3"0

3.5

T
Figure 1
F l a p p i n g response to p i t c h i n p u t - a m p l i t u d e .
T h e o r e t i c a l curves h = 0 ' 2 5 ; e x p e r i m e n t a l points h = 0.
quasi-static;
-C(k) = F + i G;
- - - C'(k, h, m) =: F" + i G'.

220 ~

200 ~

oK

180 ~

120 ~
-q~13

80 ~

40 ~

j~
0.,5

1.0

1.5

2.0 to 0 2.5

3.0

3.5

4,0

/z.5

Figure 2
F l a p p i n g response to p i t c h i n p u t - p h a s G
T h e o r e t i c a l c u r v e s h = 0"25; e x p e r i m e n t a l curves h = 0.
------quasi-static;
C(k) = F + i G ;
- C'(k, h, m) = F ' + i G .

62

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H. MOSER

1"8

"]'6

1-t~

ZAMP

I
I

t
I
I
I
I

12
1'0

I
l
l
l

\
\

0-6,
/

04

l.O

2'0

//

3 0 coz ~..0

5.0

6.0

7.0

2
Figure 3
Flapping response to displacement input at hub-amplitude.
Theoretical curves h = 0-25; experimental points h = 0.
C(k) = F + iG; - quasi-static;
c'(k, h, m) = F" + i 6";
o experimental points.

flapping response due to sinusoidal variation in hub displacement. In each


figure, the response predicted by each of the three aerodynamic theories is
plotted together with some experimental points found in the experimental part
of that study*).
The flapping responses calculated on the basis of quasi-static aerodynamics
will be recognized as essentially the responses of a second-order system with
a finite amount of damping to the different types of inputs. The effect of the
wake, as predicted by THEODORSEN[21, is to decrease the magnitudes of both
the flap-damping and the unsteady lifts, with the result that the response so
4) The authors are indebted to Messrs. JOHN ZVARA and NORXAN D. HAM, co-authors of the
report [1], which this section of the present paper summarizes in part. Figures 1, 2, 3 are reproduced
with their kind permission.

Vol. IXb, 1958

Unsteady Flow Through Helicopter Rotors

63

calculated is not very much different from that predicted by quasi-static


theory at the frequency ratios considered (although tile difference becomes
more significant as the reduced frequency increases).
However, the unsteady effects assume a new significance in vibration work
when the influence of the returning wake is taken into account. In this case
(Figures 1 and 3) the flap-damping coefficient falls off markedly as the flapping frequency ratio approaches multiples of the number of blades. It is seen
from the pitch forcing case (aerodynamic forcing) that the magnitude of the
unsteady lift also falls off at these values of frequency ratio, so that there are
'dips' in the magnitude of the flapping response to an aerodynamic input. On
the other hand, the flapping response of the articulated rotor to a displacement
input at the hub is highly sensitive to the returning wake, in that while the
driving forces are maintained essentially constant by the nature of the excitation, there is a loss of damping when the frequency ratio approaches multiple
values of the number of blades. An increased amplification ratio at these frequencies results.
It may, therefore, be concluded that the effect of the returning wake is
indeed appreciable in vibration analysis, as predicted by LOEWY and TIMMANVAN DE VOOREN, and already confirmed experimentally by DAUGHADAY,
DuWALT, and GATES. It is, in addition, noted that the two-degree-of-freedom
analyses not only predict with reasonable accuracy the effects of the returning
wake on the rotor response, but also show satisfactory correlation with the
measured responses throughout the entire range of frequency ratios considered.
Thus the importance of unsteady flow is established, and justification is provided for advancing the theory beyond the fairly severe restrictions to which
it is still subject. It would be unreasonable to infer from the above that in all
rotor vibration studies the Loewy airloads will be adequate. Both the data in
Figures 1-3 and those of [41 were taken outside of appreciable ground effect.
Furthermore, the reductions in amplitudes of airloads due to three-dimensional
flow occur in all aerodynamic terms of the equations of motion, and there is
reason to believe in the example presented here that the accuracy of twodimensional predictions of pitching and flapping response may result partly
from self-cancelling errors. One cannot assume that three-dimensional effects
will be equally secondary in every application.
3. U n s t e a d y L o a d i n g of a R o t o r in the P r e s e n c e of a G r o u n d Plane
The theory of LOEWY E3] was of course derived for rotary wings operating
in an unbounded medium. For practical reasons, however, it appears desirable
to re-cast tile problem so that the effects of a solid b o u n d a r y may be considered. This re-casting yields (1) an expression for the aerodynamic loads and
damping coefficients as a function of the distance above the ground plane,

64

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H. MOSER

ZAMP

(2) a more accurate interpretation of rotary wing test data taken on a test
pylon in the proximity of a ground plane, and (3) a rough approximation to
the aerodynamic pulse loading experienced by a rotor as it passes over a
fuselage.
The basic aerodynamic model postulated in E3] is retained, hence no attempt is made in this paper to reiterate that complete development. Furthermore, since the main consideration is a study of the ground plane effects, a
single-bladed rotor has been considered, so as to separate clearly the ground
effects from multi-blade considerations. This step promotes the interests of
clarity and simplicity, the extension to a multi-bladed rotor being straightforward.
When the basic aerodynamic model postulated by LOEWY E3] is placed in
the proximity of a ground plane, additional conditions must be fulfilled. One
obvious condition is that the velocity component normal to the ground plane
is zero along the ground plane. Also it is assumed that the shed wake does not
rebound from the ground plane. The wake is collected, held captive at the
ground level and quickly dissipated by vortex-ground plane interaction
(Figure 4). This behavior is observed, for example, when a smoke ring is
blown against a flat surface.
CO
-CO.

Cu

Shed vopticity

coptupedend
annihiI~ted by
vortex - g round
inteructton

Ground plane
Figure 4
Schematic of vortex cancellation.

The assumption of quick dissipation of the vorticity at the ground level


implies that the instant a row of shed wake touches the ground it can no longer
contribute to the downwash distribution over the airfoil. With this simple
physical picture in mind, it is easy to construct a vortex image system that
provides (1) the required normal velocity at the ground plane, (2) the condition of no wake reflection, and (3) the annihilation of all vorticity at the
ground level (Figure 5). If vortices first begin to be shed at time t = 0, the
aerodynamic model depicted in Figure 5 is valid for all times t ~ Ho/u; that
is, when at least one row of shed vorticity has contacted the ground plane.
The time history of the aerodynamic loads in the interval may be obtained by
deleting successive rows of vorticity (p, p - 1. . . . . etc.).

65

U n s t e a d y Flow T h r o u g h Helicopter Rotors

Vol. I X b , 1958

alto(x) due to image vortex

, "/"~- d ' ~ - ( l ' l ' t l ~ y SiC(lt vortex

"
n= 1

- _

n:2

/,e

n*=p ~

"7

n;2

....

n'l

,,_

(~-

....

, 2~ru

....

''

~/

g2

- ~

Ho,

Ground plane
Image vortex

"

wake system

Figure 5
P i c t u r e of theoretical model with g r o u n d plane.

Using the configuration illustrated in Figure 5 and applying the law of


BIOT and SAVART, the induced upwash (i.e., neglecting sidewash and wakewake interaction effects) on the physical airfoil resulting from the complete
vortex system is:

t) d.~' ~ra(~',
,
-~'-

2 .-,

~o(x', t)

--b'

" ~,~(r t) (.~ - r ) d~

ix~ --~')~ (/i~')"- "

--b

oo

oo

+ i ~,o!?'~-!.,
~, d ~ ' - / rA,,t)(~ - r 1 6 2
J

,~ - ~

b'

b"

(1)

W ~-~')~ + (/-h') ~"-

oo

/q

0,3

L (.;'=

(.,.'-

h':ll'

where primes (') denote dimensional quantities and

(2)

n*--2p+e-n=/-n,
/= 2p+

e.

(3)

L e t t i n g y w = -7w e'~*, and since y~o~ is the vorticity shed at some earlier time,

Now using the b o u n d a r y condition of zero pressure difference across the wake
pattern, defining a circulation function
b'

b'

/'~a d~'
Zb

ZAMP IXb/5

66

HOLT ASHLEY', t~UGENE BRUNELLE~ a n d HERBERT H. ~[OSER

ZAMP

and nondimensionalizing the upwash equation with respect to the semichord


b'; the upwash equation becomes
I

~o(~) =

" e

_ ~ - [ ~ ~-_-~ d~: - ~ k r ' / j- ~ =x - r


--

CO

d~ + ik_~

J(x-~)~+(/h)
1

P
-ik?' ZI
;~e--i2nmn--ik.~

(5)
I;

(~= ~)~ 4c- (;;~, ;;?

Ia

Resorting to integrations in the complex plane and a change of variable


x -- ~ = ~ :

L. = i ~ e -kq'- e - i ( k x + 2 . . . . ),

(6)

i:; -.= -- i ~ e -kqs e - i(kx +2nm~,)

(7)

where
q.,.= n h ,

Now letting

q3= n* h.

(~ - x) = 2, a l s o n o t i n g t h a t 9

"'~,~ e-

11 ~ - e - ' k ~ j - ~ , .

1 _< x ~ 1 a n d 0 _< ). < o 0 :

(8)

d;t.

Two more integrations in the complex plane yield the result"


il --"~- -- e - i k X [ B ( k l) -

(9)

i2-e--k2],

where
1

B(k l) = -2- { E- Ei(-- k l)l e~ - EEi(k Z)i ~-~} ,


oo

(10)

exponential integrals,
kl

Z=/h.
Collecting results, the upwash equation may be rewritten as
--

oo

~ ( x ) - 2-~-jy2_-~-d~+ ~/~
1

2~ j x - ~
--1

(11)

Vol. IXb, 1958

Unsteady FlowThrough Helicopter Rotors

67

where

(12)
n=l

Applying S6HNGEN'S [61 inversion relation to equation (11) and following the
classical work of SCHWARTZ[15], the pressure distribution is expressible as
1

-AFa(*' -- 2 [ 1 - C ' ] i V
1

V 1 + ~ 77a(~)d~

~ - x

-1

2 i(V ~

+ W

(13)

1-~

.-~

--1

where
Al(x,~)=

fin

(14)

and
c' =

Hi;'(k) + 2 Jdk) ~'?


H~-~(k) + i H~"~(k) + 2 [A(k) + i Jo(k)] I,V "

(15)

The above pressure distribution equation, has the same form as that for a
corresponding two-dimensional fixed wing, oscillating in incompressible flow.
Therefore, as noted in [31, the usual fixed-wing expressions for lift and moment
may be used when describing the simple harmonic unsteady aerodynamics of
a rotating blade section in the presence of a ground plane, provided that the
above definition of C' is substituted for the fixed-wing Theodorsen function
C(k). Additionally, the flap-damping coefficient is given by CU.D. = 2 ~ F',
and the pitch-damping coefficient E7] is given by

a~ 2 G'

where C' = F ' + i G' (a pitch axis location measured aft of the midchord in
semichords).
Therefore, by calculating IY, and hence C', all the necessary aerodynamic
information is then available to perform the standard dynamic stress analysis
routines, the 'typical section' flutter analyses and the approximate threedimensional (strip theory) flutter analyses.
Of most immediate interest, however, is an assessment of the ground-plane
effects on the rotary wing. This assessment can be made by observing the functional dependence of C', Cy.D. and Cp.D., on frequency ratio, with the wakespacing and the total rotor-ground spacing as parameters. Furthermore, since
the dependence of C', Cr.D. and Cp.D. on h' is qualitatively the same as
exhibited in [3}, only H o will be considered as a new parameter. These functional dependences are shown in Figures 6, 7, and 8. As to be expected, when

68

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H..~IOSER

r =16

1'0

ZA.MP

h'l"O0

0'8

0.8

F'
0.4

0'2

Ho, 21.33 ---,


T
Ho, co (Loewy value)

0
0

0-5

10

1.5

2-0

FPequency patio

25

3:0

3.~

4.0

m.co/f~

Figure 6
F l a p d a m p i n g coefficient.

500

~00

300

~00

I00

,~,
0~

10

)~- I-'~'~, - ~ - " ~


15

,0

25

3o

3~

,0

m
Figure 7
O--OHo=3.5556

P i t c h d a m p i n g c o e f f i c i e n t (a = 0).
; A--&
H o = 7.Illl;
....

Ho = ~

H o + oo, I~V uniformly approaches W, where W is the wake weighting function given in E3].
It can be concluded from the figures that the presence of the ground plane
provides additional positive aerodynamic damping to all blade modes of

69

Unsteady Flow Through Helicopter Rotors

Vol. I X b , 1958

80

60

20

Y
-~0

0"5

l'O

1.5

2"0

2'5

3.0

3.5

4,0

m
1;igure 8
P i t c h d a m p i n g coefficient (a = -- 1).
C P, D. vPI'S|lS II$.

motion, hence greatly reducing the possibility for one-degree-of-freedom


flutter to occur when the helicopter rotor plane is less than one rotor diameter
above the ground plane. Note, however, that this type of flutter is still possible
under unfavorable circumstances. When the helicopter rotor plane is more
than two rotor diameters above the ground plane, the ground effects are
completely negligible and t'V ~- W.

4. T h r e e - D i m e n s i o n a l

U n s t e a d y F l o w Effects

Flow past a single helicopter rotor blade is three-dimensional because of


the curvature of the streamlines, the presence of 'wingtips' and the spanwise
variation of relative wind. Since the blade aspect ratio is usually quite large,
the third effect may have equal or greater importance than the other two. As
applied to rotors, two-dimensional strip theory contains an inconsistency, associated with this nonuniform relative wind, which does not occur in the case
of a conventional airplane wing: even in steady rotation at constant angle o f
attack the bound circulation varies linearly with radial distance from the hub
thus violating the law of continuity of vortex lines. For steady, vertical flight
of a lightly-loaded rotor this and other objections are fully overcome by GOLDSTEIN'S vortex model [8]. Discussions of several other approximate, steadystate theories of the finite rotor will be found in NIKOLSKY [9~. An elaborate

70

I-IOLTASHLEY, EUGENE BRUNELLE~ and HERBERT H. ~IOSER

ZAMP

vortex representation of the hovering condition has been given recently by


GRAY and NIKQLSKY [10].
Sections 1 and 2 above demonstrate the need for a theory that predicts
loads due to unsteady, three-dimensional flow. Since such a theory would perforce be linearized and its results therefore capable of superposition, a restriction to simple harmonic motion is acceptable. It is not surprising, however,
that no efforts along these lines appear to have been published; no fully
rational theory exists even for the finite airplane wing in unsteady motion, and
most of the suggested approximations involve severe mathematical complexities. The paragraphs below describe a first attempt at analyzing the oscillating
finite rotor. The authors freely admit that radical modifications of the true
physical system have been made in constructing their vortex model. Even then,
the computations remain lengthy, and only a few examples have been worked
out. Two facts lent encouragement in carrying the development through to
completion. First, the large aspect ratios of typical rotors insure that a theory
of the lifting-line type will furnish the loads needed for nearly any practical
dynamic calculation. Second, a modified lifting-line theory was found to yield
satisfactory steady-flow results.
To construct a steady lifting-line theory for vertical flight of a blade whose
root and tip lie at r = R1 and r - - R 2, respectively, let the bound circulation
/'(r) be concentrated in a single straight vortex along the quarter-chord line.
As seen from a coordinate system mounted on the blade, the oncoming stream
has a velocity V(r) = .(2 r normal to the lifting line. The vortices trailing into
the rotor wake, whose circulation is dr'/dr per unit radial distance, are assumed
to extend straight downstream to infinity. Hence one neglects both the helical
shape of the actual wake and the centrifugal terms in the equations of fluid
motion relative to the rotating coordinates, approximations which can be
justified under certain circumstances but which are best assessed from the accuracy of the results. One is thus led to the problem of an ordinary airplane
wing in a sheared main stream, and by direct analogy with conventional
lifting-line theory (cf. GLAUERT [11], chaps. 10-11), the bound circulation
satisfies the integrodifferential equation
R~

Cla b

s~

4~9r

r-,l

RI

Here cl~, b', and s o denote the local lift-curve slope, blade semichord, and
incidence measured from zero-lift, all of which may be functions of r. ~Vhen
equation (16) is solved by well-known methods, it yields spanwise load distributions and rotor thrust coefficients which fall among the values estimated b y
more commonly used theoretical methods Eg]. (Certain other quantities such

Vol. IXb, 1958

Unsteady Flow Through Helicopter Rotors

71

as induced inflow velocity do not compare so favorably, but the emphasis here
is on airloads.)
The theory of the oscillating finite wing in incompressible flow which is a
direct extension of lifting-line theory, and which reduces thereto when the frequency approaches zero, is REISSNER'S E12]. The authors have, in effect,
modified REISSNER'S physical approximations and mathematical derivation to
include the effect of a space-variable main stream V(r). It is impossible to
present the detailed steps here. Only the general outline is given, but the
interested reader can find the full development in [13].

Figure 9
A rotor blade osciUating sinusoidally in a sheared flow, showing locatkms of coordinate systems.
Definitions:
1

r'

,'=,--Ro;

+7'

~*"~*= b~' -hi-;

(z', ~') - o- [zi + zl]

(z,~-)=

z; + z~

~=

~b--~o"

-~ [z; - zl]

Figure 9 illustrates the rotor blade in a sheared flow and defines the dimensional and dimensionless coordinates used in the development. The motion
of the blade is specified in terms of the normal velocity component of fluid
particles in contact with it (upwash)
va(x', r', t) = -~,~ (x', r') e i''t ,

(17)

which is assumed to be a known quantity. The flow over the blade and its wake
is approximated b y a vortex sheet in the (x' - r')-plane with circulation components (velocity discontinuities) 7 spanwise and ~ chordwise. The amplitude
of total circulation ['(r') bound to the blade is replaced by a modified circulation function/w, defined, following REISSXER [12!, as
]~(r') = r ( 7 ) e.,,~,v
b~

(18)

Application of the Biot-Savart law, the K n t t a condition of smooth flow-off

from the trailing edge and the requirement of zero pressure discontinuity

72

HOLT ASHLEY, EUGENE BRUNELLE, a n d HERBERT H . MOSER

ZAMP

through the wake vortex sheet leads to the lifting-surface integral equation
(factor ei~ has been cancelled)

_ ,
vo (x, r') =

~
4~

sb~ x t _

~,~(~,,~)[x ~ r
(~- = ~U#

~ = 3~3 ~

-,/]

a~' d,/

--sb'. x t

b6
4~

--sb'. x t

d [p,(,) e_~e,,v (~'1]It" - ~'] - ~ ~


P ( , / ) e -~'~'zl"'~ Ix" - ~'~
dn'
{Ix' - ~']~ + fr' - V']~p: ~-

d~' d~'.[ (19)

Equation (19) involves only the approximations of ordinary lifting-surface


theory and is the counterpart of equation (18), Part I, of REISSNER [12].
Reduction of equation (19) to a form suitable for convenient mathematical
manipulation requires several modifications of the physical model, two of
which have direct parallels in the development of lifting-line theory. These
may be stated in words as follows:
(1) The bound vortices are treated as if the blade were replaced by an airfoil
in two-dimensional flow, having the same chordwise toad distribution as
the section at station r' ~i.e., 5n ~- 0 and 7-~ (~', 7') -~ ~ (~o, r') in the first
integralJ.
(2) The trailing vortex pattern in the wake is projected forward from the
trailing edge xt(~f) to a spanwise line passing through point (x', r'), where
tile vertical velocity is to be determined.
(3) In the pattern of shed 7,o-vortices, those portions which represent deviations from two-dimensional flow at any spanwise station are also projected up to a spanwise line through (x', r'). In this same pattern the
wavelength factor e-~r
is approximated by e- i,o~'i7(/)
When the foregoing steps are carried out and the dimensionless notation
of Figure 9 is introduced into equation (19), one obtains
1

~o(z,

y*)

....

c'~

L. d:~(r162

+ i ko T'(~*) ~-"~ .... 1" ~ - ~

--1

dr

4 . y at/* k~

.TVB (ko(~*) Jr* -- ~/'3)

--s

+ i ~-' t~~
"+ 4~'z

('rl*) ~

t V(t/*)]

:Vo (ko ('*) E'* -- ~*j)} d~*


e-i[/~ . . . .

+/:zit'(,*)/v(,~*)

--$

V(r*)

(N~ (ko(,~*/I~*-~,~)+ Eke+ kO~m?V--r N, (ko('~*/E~*--~*l))


-

~-~ .....+~J ~? (ko(r*)?*

- ~*~)~/ dr*.

(2o)

Vol. IXb,

Unsteady

1958

73

Flow Through Helicopter Rotors

Here the reduced frequency appears in four forms


k(~*) -

o~ v ( ~ * )

v(~*)

'

ko(r,)_ v(~*)
oJ b6 '

o~ b'(n*)

o~bd

k(,~*) k0(~*) -

y (n*) '

the a r g u m e n t r* being implied wherever it is not indicated.


Three special functions arise in this development which are related to the
so-called Cicala function of u n s t e a d y wing theory El21. Unlike the latter, two
of t h e m can be evaluated in terms of integral representations of t a b u l a t e d
functions, as given b y WATSON" [14]. The definitions and evaluations are listed
below :
oo

..

~q
/" Ze-"d3.
[ql [1 - - i q K 0 ([q])] + -2- [L~
NA (q) = q J [fi-+ }~i5~-~ -- q

--

I~

(22)

2v)~(q) = q j

]"

e - z ) " d).
1
-j~-~]_~-~
. . . . q- + AN~(q)

(23)

= 1 _ i [q/ + ~
[/l(q) - Ll(q)] +
q
q
-

I ] Kl(!q] ) _

q '

oo

Nc(q) =

-)3." q;
o

q I-D-- d2
l~p

= l qL In 1 2 q ] +
q

[L0(x ) - I o ( x ) ] d x - i

(24)

!q:

Ko(X)dx

Here I~ and Kn are modified Bessel functions, while L~ is the modified Struve
function [14] of the first kind and order n. The pole singularity of N8 at q = 0,
which is just the steady-state singularity of lifting-line theory, has been separated for c o m p u t a t i o n a l convenience, leaving a remainder ANB which is everywhere finite. Na and Nc have no singularities for real, finite arguments excepting an integrable, logarithmic singularity of Nc at the origin. No particular difficulty was encountered when constructing tables of these functions
over the ranges needed, one convenience being t h a t they are all odd.
The integral equation (20) is readily solved for ~ b y well-known methods
of thin airfoil theory (cf. SCH~NARZ [153). The result can be cast as the s u m of
two-dimensional theory plus a three-dimensional correction c o n t a i n i n g / " and
the N-functions, In view of the relation
xt

/ r o ( ~ , y') d x ' = b / to(z, ~*) d z =

xz

--

"'

'

--I

--

b'o e - i [ k + k ~

?'(r*)

(25)

74

HOLT ASHLEY', EUGENE BRUNELLE, a n d HERBERT I-{. MOSER'

ZAMP

the expression for Ya can be integrated chordwise to produce an integrodifferential equation for _~', which is the principal unknown
F '(~)= F ' + ;r i [Ht~(k) + i H~J(k)]

x NB (ko(,*) ~r* -

7*]) +

e-ik~

if'(~*)~,

V(~*) : -- i f :

~V~-*y! Na

V01*)

(ko(rl*) It* - 7*])]

@*
s

- r ~(k)-;I.(k)l,-'~

-- d~l*

....

f-~,, [F(:) ~((~-,)-j


v(,.)] N c (ko(,*) Jr* -- 7"1) d~* (26)

--S

+ i ko

v(.*), d.,.
d iv(.*)]
/~(~*) -P(~-~y
t v(,:)j N. (ko(~*) [,* - 7*]) ' - " : " 7(7*)
v(,*)

--8

_V(r*) ~

(.o[fo(,,

V(r*) ,1

J,

[:o

Here/~,('-'1 denotes the circulation function as given by two-dimensional methods


[2],

b'
~'(~) = 4 ....
b6 eik~

~/-i--~-o(r162
(27)

~-:i k[H~'-'~(k) + i H[;2~(k)]

REISSNER'S linearized relation [12],


dp.(z, r*)
---

~V- ......

(z, r*) + i k / ~

~a

(28)

(~, r*) d ;
a

-1

between Ya and the complex amplitude of pressure difference between the


upper and lower surfaces of the rotor blade can be carried over into the present
theory. When equation (28) is applied to the inverted form of equation (20),
one obtains after considerable manipulation
1

,fo(....)

. I:

r( )l f

;.l:. ,.)

--1

+Wf[~ZVI-r

(z-r

z'~-~lT~'#T-fi/j (29,

--I

I~- z [Fo(r, ) + z F:(r*) + z ~ G(r*)]


-ff~(~, r*) d~" -- V -S+~-

VoL IXb, 1958

Unsteady Flow Through Helicopter Rotors

75

Here the first two lines are the strip-theory result, and Fo, F~, F~ represent
rather elaborate integrals across the rotor span, containing ~ ' and the N-functions but independent of the chordwise coordinate. Suitable chordwise integrations of Ap-~ give the complex amplitudes of lift force (positive upward)
per unit span and pitching moment (positive nose-up about a midchord axis)
per unit span'
E
~"
[ Fo(r*) + 1
Fl(r* )
1
F2(r*)]
q V~b' -- qv"-b' + ~ -- - - V - -2- " - - ( 1 - - -- ~ "
V .'

(30)

-FI
~ l '2'
~ [ Fo(r* )
Fl(r* )
3
F2(r*) ]
e V2 b'2 -- e V2 b'2 + - 2 - -- ~
+
V
---4- "
V ]"

(31)

The practical application of the foregoing theory is carried out in two steps.
First, the circulation integral equation is solved by introducing the angle
variables
r* = s cos(F, ~* = s cos 0,
(32)
approximating the circulation function by the Fourier series
I?'(r *) ~= ~

(33)

K~ si)~n_ ~

and requiring equation (26) to be satisfied at as many spanwise stations (tips


excluded) as there are constants K,, to be determined. Second, the loadings are
calculated by substituting equation (33) into the expressions for P0, /71, F 2,
then applying equations (30) and (31). This process resembles closely that
devised by REISSNER and STEVENS [12], Part II. In the first examples, however, it was found convenient to use series expansions of f0, J1 and the complex
exponentials; these were carried up to terms of orders k 2 and ko, which appears adequate for the majority of helicopter applications. It proved necessary
to evaluate numerically several hundred integrals containing the N-functions,
sinn 0, cosn 0 and various powers of
V(r*)
v(~)-

R*/s + cos ~

R,* + r*
=

-R~;-+ ~*

R*/~ +

(34)

Co~0

as factors in their integrands. Although the first examples were laborious, the formulation is now systematized and well adapted to high-speed digital computation.
To obtain some idea of the importance of three-dimensional effects on
vibrating rotors, the theory has been used to compute the loads due to pitching
and flapping of a constant-chord blade with aspect ratio 8. The frequency is
given by r
= 1. (Since the relative wind is everywhere proportional to rotor
speed s it is unnecessary to specify a reference value of reduced frequency k ;
all dimensionless quantities are fixed by c~/Y2LFive spanwise stations were
chosen, spaced equally in the variable 9.) Figure 10 plots, vs. radial distance,

76

HOLT ASHLEY, EUGENE BRUNELLE, a n d HERBERT H. 3,lOSER

2~~

'

~,/

TnPee-Oimensionat

phase

/ 2 o

Two-rtimensionnt

am0tituoe

- --7"

..,~3

/
~

ZAMP

"x,

0.2

0-t

/...~--._Tm'ee-tlimenslonall

~
160

12" ~,
8" ==

phase

"1

0"6

0"8

1-04 ~

It2
Figure l0
Dimensionless a m p l i t u d e and phase angle of lift due to pitching oscillation, a b o u t the q u a r t e r - c h o r d
axis for a c o n s t a n t - c h o r d r o t o r with R 1 = 8 in., Ro = 40 in., 2 b = 5 in. Circular f r e q u e n c y ~o = .O.

the dimensionless amplitude and phase angle of lift per unit span due to nose-up
pitching oscillation
c~ = ~ e i~,t
(35)
about an axis along the quarter-chord line. Dimensions of the blade planform
are listed in the caption. ~pr~ is the phase angle b y which the force leads the
angular displacement; on a quasi-steady basis it would equal zero. The curves
marked ' two-dimensional' represent the unsteady strip theory of THEODORSEa"
[2], which is discussed in previous sections. Other airloads due to pitching and
flapping show the same general behavior as Figure 10, for the particular rotor
and value of o)/.(2 chosen here.
It can be concluded that in this example the phase shift produced b y threedimensionality of the flow is small enough to be negligible when one is analyzing forced motion. The reduction of amplitude is more significant, particularly
in the tip region. The total lift drops roughly 2 5 0 , while the moment about the
flapping hinge of the blade is reduced by well over 30%. Further remarks on
the imphcations of this calculation are made in the next section.
The three-dimensional theory can be modified in an approximate way to
account either for the returning shed vortices in the helical wake during vertical flight, or for the small fluctuations of relative wind due to forward flight
at low tip-speed ratios. It is unlikely that both of these effects would have to
be introduced simultaneously.

Vol. IXb, 1958

Unsteady Flow Through Helicopter Rotors

77

Without going into detail, the shed-vortex modification is made on a twodimensional basis b y analogy with LOEWY'S work [31. The two-dimensional
circulation function of equations (27) and (26) must be replaced by (midchord
sweep zm is set equal to zero here for simplicity)
1

2 Jfl/
V 1 - ~ ~(~' r*) d~
Ft(2) =

--1

j%
} .
~ i k t 2 [ni2}(k) + i H~-~(k)]+ [ y # ) + iYo(k)] W

(36)

Here W is a wake-correction function [31, which approaches zero as tile spacing


between successive helical vortex sheets becomes large. The circulation integral
equation is then solved as before, but in the two-dimensional parts of the final
pressure, force and moment formulas the function C(k) must everywhere be
replaced with the function C'(k, m, h') tabulated by LOEWY. This procedure
omits the potentially large influence of the strong helical vortices trailing from
the neighborhood of the blade tips, something which certainly merits further
study but will lead to serious analytical complications.
Forward speed can be accounted for two-dimensionally by reference to the
work of GREENBERG [16i, who studies an airfoil operating in a stream whose
speed is given by the real part of
V = Vo [l + a ei~"t 2 .

(37)

Some care must be observed when handling terms containing the combined
factors e i~'t and e i ' , t , but for a particular motion like flapping or pitching,
the principal changes come in the function F'(") and the two-dimensional parts
of the loading expressions. It is important to realize that GREENBER~'S theory
contains the implicit restriction that a / k , is small compared with unity, where
k, = co, b ' / V o. This may limit its applicability to typical rotor problems rather
severely.
No numerical examples involving forward speed or shed vortices have been
worked out as yet. From the general formulation, however, there is good
reason to believe that three-dimensional loadings incorporating the latter will
exhibit the same drops in magnitude and losses of aerodynamic damping near
integral values of m that were discovered in LOEWY'S original investigation.
5. C o n c l u s i o n s

Although much remains to be done in the way of practical application of


the theories described in sections 3 and 4, some tentative conclusions can now
be drawn concerning rotor vibrations and unsteady flow. When calculating
forced vibratory response during hovering or vertical flight, unsteady effects

78

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H. MOSER

ZAMZ'

must be accounted for, with particular reference to the shed vortices in the
helical wake. If this is not done, potentially dangerous resonances near frequency ratios which are integral multiples of the number of blades wilt be overlooked. It can be inferred that aeroelastic stability is subject to the same influences, and that rotor flutter speeds much lower than those computed without regard for the returning wake may occur whenever the flutter frequency is
close to any critical multiple of Q.
In the presence of the ground, the dangers of resonant vibration and flutter
are reduced, since the ground plane provides additional aerodynamic damping
for all modes of blade motion. Single-degree-of-freedom flutter can still occur
in an unfavorable combination of circumstances. When the rotor plane is more
than two diameters above the surface, ground effects become negligible.
For values of the reduced oscillation frequency typical of helicopter practice, the three-dimensionality of the flow does not cause significant changes in
the phase angle between a given motion and the airloads produced thereby.
There will be a large amplitude reduction, however, especially near the rotor
tip. This suggests, incidentally, that finite-span effects can perhaps be estimated on a steady-state basis. It cannot be concluded that forced vibratory
responses calculated by strip theory will always be larger than those actually
encountered, because three-dimensional effects reduce both the forcing functions and the aerodynamic 'springs' and 'dampers' opposing them. Since
these reductions are not the same for all terms in the equations of motion,
further study is needed to discover the conditions under which strip-theory
predictions are unacceptable. Three-dimensional flow is likely to be more
significant for flutter analyses, since flutter occurs at a stability boundary,
whose location may be sensitive to small changes in the aerodynamic properties
of the system.
Tile present investigation concentrates on the case of vertical flight. It is
possible, however, to adjust both conventional strip theory and the threedimensional approach of section 4 to account for forward motion at small tipspeed ratios. This can be done along the lines proposed by GREENBERG [t6].
Since the most severe forced vibrations are often encountered on helicopters
in this transitional range, further study of the problem is very desirable.

REFERENCES
Eli j. ZVARA, N. D. HA.~I und H. H. MOSER, E[[ects of Unsteady Aerodynamics on
Helicopter Rotors, Part I, to be issued as a "W.A.D.C. Technical Report (1958).
~2] T. THEOOORSEN,General Theory o/Aerodynamic Instability and the Mechanism
o/Flutter, Report 496, National Advisory Committee for Aeronautics (1935).
[3] R. G. LoEwY, A Two-Dimensional Approximation to the Unsteady Aerodynamics o/Rotary Wings, Report 75, Cornell Aeronautical Laboratory (1955).

Vol. IXb, 1958

Unsteady Flow Through Helicopter Rotors

79

[4J H. DAUGHADAY,F. DUWALDT,and C. GATES, Investigation o/Helicopter Blade


Flutter and Load Amplification Problems, Institute of the Aeronautical Sciences,
Preprint 705 (1957).
[SJ R. TIMMANand A. I. VAN DE VOORE~r Flutter of a Helicopter Rotor Rotating in
Its Own Wake, J. aeron. Sci. 24, No. 9 (1957).
[6j H. S6HI~GEI%Die L6sungen der Integralgleichung
a

g(x) = ~

yd ~/(~)

dk

und deren Anwendung in der Tragfli~geltheorie, Math. Z. 45, 245-264 (1939),


E7] H. L. RUNYAN, Single-Degree-o/-Freedom Flutter Calculations /or a Wing in
Subsonic Potential Flow and Comparison with an Experiment, Technical Note
2396, National Advisory Committee for Aeronautics (1951).
[8J S. GOLDSTEIN, On the Vortex Theory o/ Screw Propellers, Proc. roy. Soc. [AJ
723, 440-465 (1929).
[gJ A. A. NIKOLSKY, Helicopter Analysis (John Wiley & Sons, New York 1951).
El0] A. A. NIKOLSKYand R. B. GRAY, On the Motion o/the Helical Vortex Shed/rom
a Single-Bladed Hovering Model Helicopter Rotor and Its Application to the Calculation o/the Spanwise Aerodynamic Loading, Princeton University Aeronautical Engineering Dept. Rept. No. 313 (Sept. 1955).
[11] H. GLAUERT, Aero/oil and Airscrew Theory, American edition (The Macmillan
Co., New York 1943).
[12J E. REISSNER, E//ect o/ Finite Span on the Airload Distributions/or Oscillating
Wings, Parts I and II, Technical Notes 1194 and 1195, National Advisory
Committee for Aeronautics (1947).
[13J H. ASHLEY, J. ])UGUNDJI, and H.H. MOSER, E//ects o/ Unsteady Aerodynamics
on Helicopter Rotors, Part I[I, to be issued as a W.A.D.C. Technical Report
(1958).
[14] G. N. WATSON, Treatise on the Theory o/ Bessel Functions, Second edition
(The MacMillan Co., New York 1945).
[15J L. SCHWARZ,Berechnung der Druckverteilung einer harmonisch sich ver]ormenden
Tragfldche in ebener Str6mung, Luftfahrtforschung (17. Dez. 1940), p. 379-386.
I16] J. M. GREENBERG, Air/oil in Sinusoidal Motion in a Pulsating Stream, Technical Note 1326, National Advisory Committee for Aeronautics (1947).
Zu sam m en/ass vlng
Es wird ein 13eitrag geliefert zur Stfitzung der schon frfiher ge~usserten ~)berzeugung, dass nichtstationXre Str6mungsanteile das Schwingungsverhalten und die
aeroelastische Stabilit~t yon Helikopter-Drehfliigeln wesentlich beeinflussen k6nnen. Die Untersuchung betont besonders den D~mpfungsverlust, welcher sich aus
den entlang der Spannweite eines schwebenden Drehfliigels schraubenf6rmig abgehenden Wirbeln ergibt.
Gem~ss dem Bediirfnis ftir verbesserte Mitre1 zur Behandlung nichtstation~rer
aerodynamischer Vorg~nge werden zwei neue theoretische Entwicklungen beschrieben und zum Berechnen yon Luftkr~ften auf typische schwingende Drehflfigel
angewendet.
Die erste dieser Arbeiten erfasst auf zweidimensionaler Basis das Vorhandensein
einer Bodenebene. Es wird angenommen, class die \Virbel im Abwinde des schwe-

80

HOLT ASHLEY, EUGENE BRUNELLE, and HERBERT H..~[OSER

ZAMP

benden Drehfliigels durch die Bodenebene nicht zuriickgeworfen, sondern schnell


zerstreut werden. Als Resultat zeigt sich, dass die einer spezifischen Schwingungsform entsprechende aerodynamische D~impfung durch dell Einfluss des Bodens,
der praktisch etwa zwei Drehfliigel-Durchmesser hinaufreicht, im allgemeinen vergr6ssert wird.
Die zweite der vorgelegten Theorien betrifft die dreidilnensionale Str6mung
/iber ein schwingendes Fliigelblatt, welches als tragende Linie und unter Vernachliissigung der Kriimmung seines Abwindes untersucht wird. Numerische Rechnungen sind ziemlich miihsam, aber mittels Rechenmaschinen ohne weiteres durchfiihrbar. I n einem herausgegriffenen typischen Beispiele zeigte sich, dass die Luftkriifte
bedeutend kleiner sind als die yon der Streifentheorie angezeigten, dass aber die
Phasendifferenz zwischen Anstellwinkel und Luftkr~ften fast unveriindert bleibt.
(Received: October 3, 1957.)

You might also like