You are on page 1of 51

Aeromechanics of Rotary Wing

Vehicles
Class Notes for AE 545

Peretz P. Friedmann

FXB Professor

Department of Aerospace Engineering

University of Michigan

September 13, 2010


CONTENTS

1 Introduction to Rotary Wing Aerodynamics 1

1.1 Momentum Theory of Rotors . . . . . . . . . . . . . . . . . . 1

2 Blade Element Theory 5

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.2 Rotor Torque . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.3 Vertical Climb . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.4 Rotor Tip Losses . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.5 Extended Blade Element Theory . . . . . . . . . . . . . . . . 19

2.5.1 Optimum Rotor for Hover . . . . . . . . . . . . . . . . 22

2.5.2 Rotors Hovering in Ground Effect . . . . . . . . . . . . 23

2.5.3 Descent in Axial Flight . . . . . . . . . . . . . . . . . . 27

2.6 Blade Element Theory of Descent Problem . . . . . . . . . . . 34

2.6.1 Partial Powered Vertical Descent . . . . . . . . . . . . 36

2.7 Autorotation . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.8 Rotors in Non-axial Forward Flight - or Momentum Analysis

in Forward Flight . . . . . . . . . . . . . . . . . . . . . . . . . 46

2.9 Rotors in Non-axial Forward Flight (Alternate Derivation) . . 48

ii
iii

2.10 Simple Control Concepts for Helicopters . . . . . . . . . . . . 54

2.11 Response of a Fully Articulated Rotor with Centrally Hinged

Blades to Cyclic Pitch Control . . . . . . . . . . . . . . . . . . 56

2.12 Blade Element Theory in Non-axial Flight and Some Forward

Flight Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

2.12.1 Flapping Motion and Flapping Angles . . . . . . . . . 65

2.12.2 Reversed Flow Region and Stall . . . . . . . . . . . . . 67

2.12.3 Torque Coefficients . . . . . . . . . . . . . . . . . . . . 69

2.12.4 Horizontal Force . . . . . . . . . . . . . . . . . . . . . 70

2.12.5 Lateral Force . . . . . . . . . . . . . . . . . . . . . . . 71

2.13 Forward Flight Performance Using Blade Element Theory . . 72

2.14 Forward Flight Performance Using Energy Concepts . . . . . . 74

2.14.1 Induced Power . . . . . . . . . . . . . . . . . . . . . . 74

2.14.2 Profile Drag Power . . . . . . . . . . . . . . . . . . . . 74

2.14.3 Parasite Power . . . . . . . . . . . . . . . . . . . . . . 78

2.14.4 Climb Power . . . . . . . . . . . . . . . . . . . . . . . 79

2.14.5 Total Power . . . . . . . . . . . . . . . . . . . . . . . . 80

2.15 Reference Planes Used in Forward Flight Analysis . . . . . . . 81

2.15.1 Change in Reference Planes . . . . . . . . . . . . . . . 83

3 Flapping Dynamics of a Rotor Blade 86

3.1 Forward Flight Case . . . . . . . . . . . . . . . . . . . . . . . 86


iv

3.2 Special Case for Hovering Flight . . . . . . . . . . . . . . . . . 98

3.3 Spacial Case for Forward Flight . . . . . . . . . . . . . . . . . 100

3.4 Helicopter Control Moments . . . . . . . . . . . . . . . . . . . 101

3.5 Application to the Hingeless Rotor . . . . . . . . . . . . . . . 102

3.6 A More Comprehensive Treatment of Blade Flapping Dynamics104

3.6.1 Aerodynamic Moment in Hover Case . . . . . . . . . . 106

3.6.2 Equations of Motion in Hover . . . . . . . . . . . . . . 108

3.6.3 Response to Cyclic Control in Hover . . . . . . . . . . 111

3.6.4 Blade Dynamics in Forward Flight for Stability and

Control Applications . . . . . . . . . . . . . . . . . . . 112

4 Trim Analysis 122

4.1 Propulsive Trim . . . . . . . . . . . . . . . . . . . . . . . . . . 123

4.2 Some Additional Comments on the Solution of the Trim Equa-

tions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

5 “Classical” Flutter of Rotor Blades 136

5.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . 136

5.2 Stability Boundaries . . . . . . . . . . . . . . . . . . . . . . . 146

5.2.1 Static Stability-Divergence . . . . . . . . . . . . . . . . 147

5.2.2 Dynamic Stability- Flutter . . . . . . . . . . . . . . . . 148

5.3 A Brief Critique of the Classical Approach . . . . . . . . . . . 156


v

6 Dynamic Equations of Equilibrium of a Centrally Hinged

Spring Restrained Blade in Hover with Coupled Flap, Lag,

and Torsional Dynamics 157

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

6.1 Coordinate Systems and Ordering Schemes . . . . . . . . . . . 165

6.2 Distributed Inertia Loads on the Blade . . . . . . . . . . . . . 167

6.3 Distributed Aerodynamic Loads . . . . . . . . . . . . . . . . . 176

6.4 Distributed Structural Damping Loads . . . . . . . . . . . . . 185

6.5 Rotor Blade Equations . . . . . . . . . . . . . . . . . . . . . . 186

6.6 Coupled Flap-Lag Stability of Hingeless Rotor Blades . . . . . 193

6.6.1 Equations of Dynamic Equilibrium . . . . . . . . . . . 193

6.7 Solution of the Coupled Flap-Lag Problem . . . . . . . . . . . 199

6.7.1 Flutter Boundaries . . . . . . . . . . . . . . . . . . . . 205

7 Some Typical Numerical Results 210

7.1 General Comments . . . . . . . . . . . . . . . . . . . . . . . . 210

7.2 Numerical Results and Comments . . . . . . . . . . . . . . . . 210

8 Effect of Additional Degrees of Freedom on the Stability of

Rotor Blades in Hover 226

8.1 Coupled Flap-Lag-Torsional Behavior of Blades in Hover . . . 226

8.2 Coupled Flap-Lag-Torsional Results in Hover . . . . . . . . . . 230


vi

References 237

Appendix 240
CHAPTER 1

Introduction to Rotary Wing Aerodynamics

1.1 Momentum Theory of Rotors (Actuator Disk Theory)

Assumptions:

1. Low disk loading

2. No rotational effects

3. Incompressible flow

4. Assume v-inflow velocity, uniformly distributed over the disk (Equiva-

lent to assuming infinite number of blades)

5. Hover

Mass flow through rotor = ρAv = ṁ (1.1.1)

Thrust from momentum equation is given by considering stations 0 and 3,

shown in Figure 1.1.

T = ρAv (w − 0) = ρAvw = ṁw (1.1.2)

1
c
Copyright Peretz P. Friedmann

Figure 1.1: Rotor slipstream

Applying Bernoulli’s equation between stations 0 and 1

ρ
p0 = p∞ = p1 + v 2 (1.1.3)
2

Applying Bernoulli’s equation between stations 2 and 3

ρ ρ
p2 + v 2 = p0 + w2
2 2

ρ 2 
p2 − p0 = w − v2 (1.1.4)
2

From Eq. (1.1.3) and (1.1.4)

2
c
Copyright Peretz P. Friedmann

ρ 2  ρ ρ
p2 − p1 = ∆P = p0 + w − v 2 − p0 + v 2 = w2 (1.1.5)
2 2 2

From Eq. (1.1.2) and (1.1.5)

ρ 2
T = ∆P A = (p2 − p1 ) A = w A = ρAvw (1.1.6)
2

From Eq. (1.1.2) and (1.1.6)

T = 2ρAv 2 (1.1.7)

or, the inflow is given by

s
T
v= (1.1.8)
2ρA

The inflow ratio is defined as,

s r
v T CT
λ= = = (1.1.9)
ΩR 2ρAΩ2 R2 2

where,

T
CT = is the thrust co-efficient of the rotor (1.1.10)
ρAΩ2R2

The dynamic pressure variation in the axial direction is shown in Fig. 1.2.

3
c
Copyright Peretz P. Friedmann

Figure 1.2: Pressure variation plot in axial direction

4
CHAPTER 2

Blade Element Theory

2.1 Introduction

Consider the case of hovering flight or axial flight:

Assumptions:

1. Blade chord c = const., pitch setting θ0 = const.

2. Axial flight (radial flow effects are negligible)

3. Incompressible flow

4. Low disk loading

Denote the velocity of vertical flight by V .

From Fig. 2.1, the blade element angle of attack is given by,

 
−1 V +v
α (r) = θ0 − tan (2.1.1)
Ωr

For low axial velocities, V + v << Ωr, and a radial position sufficiently

outboard from the axis of rotation.

From Fig. 2.1, the blade element angle of attack is given by,
 
V +v
α (r) ∼
= θ0 − (2.1.2)
Ωr

5
c
Copyright Peretz P. Friedmann

The elemental lift associated with a segment of width dr is,

1   1
dL = ρ (Ωr)2 + (V + v)2 Cl cdr ∼ 2
= ρ (Ωr) Cl cdr (2.1.3)
2 2

where, the following assumption is used:

 2
V +v
<< 1
Ωr

Figure 2.1: Geometry for blade element theory

6
c
Copyright Peretz P. Friedmann

Recall

Cl = aα (r) where a is the lift curve slope

Therefore

  
1 2 1 2 V +v
dL = ρ (Ωr) aα (r) cdr = ρ (Ωr) ac θ0 − dr (2.1.4)
2 2 Ωr

Similarly one can write the following expression for drag

1
dD0 = ρ (Ωr)2 Cd0 cdr (2.1.5)
2

where, Cd0 is assumed to be some representative constant profile drag coef-

ficient, which in reality can be considered to be of the form,

2
Cd0 = δ0 + δ1α (r) + δ2 α (r) (2.1.6)

where,

δ0, δ1 , δ2 are some constants which can be determined.

In rotor dynamics it is customary to resolve things into components per-

pendicular and parallel to the plane of rotation (hub plane).

The inflow angle is given by,

   
−1 V +v ∼ V +v
φ (r) = tan = using small angle assumptions
Ωr Ωr
(2.1.7)

Referring back to Fig. 2.1 and taking components,

7
c
Copyright Peretz P. Friedmann

CD0

α
Usual operating range

Figure 2.2: Profile drag coefficient as a function of effective angle of attack

dT = dL cos φ (r) − dD0 sin φ (r) (2.1.8)


| {z }
negligible

dD0 = dL sin φ (r) + dD0 cos φ (r) (2.1.9)

The lift L is approximately one order of magnitude larger than the drag
 
L ∼
D = 6 − 9 and thus the last term in Eqn. 2.1.8 is negligible.
D
Using small angle approximation for φ (r),

cos φ (r) ∼
= 1.0; sin φ (r) ∼
= φ (r)

Thus,

dT ∼
= dL (2.1.10)

dD0 = φ (r) dL +dD0 (2.1.11)


| {z }
Induced drag

8
c
Copyright Peretz P. Friedmann

For a rotor with b blades, the total thrust is given by using 2.1.10 and
r
2.1.11 with r = as integration variable,
R

Z R Z 1   
1 V +v
T =b dT = ρΩ2 R3 abc 2
r θ0 − dr
0 2 0 ΩRr

  
1 2 3 θ0 V +v
= ρΩ R abc − (2.1.12)
2 3 2ΩR

Recall
T
CT =
ρ (πR2) Ω2 R2

Define
Blade area bcR bc
Solidity = σ = = 2
=
Disk area πR πR

Inflow ratio for axial flight is,

V +v
λ= (2.1.13)
ΩR

Thus Eq. (2.1.12) can be rewritten as,

 
σa θ0 λ
CT = − (2.1.14)
2 3 2

2.2 Rotor Torque

Another important quantity associated with rotor behavior is the torque,

or the moment needed to overcome the drag, and keep the rotor turning at

a certain RPM in steady state conditions.

9
c
Copyright Peretz P. Friedmann

One can express the elemental torque as,

dQ = rdD0 dr

and the total torque as


Z R Z R
0
Q=b rdD dr = b (dLφ + dD0 ) rdr
0 0

Z R       
1 2 2 V +v V +v 1 2 2
=b r ρΩ r ac θ0 − + ρcCd0 Ω r dr
0 2 Ωr Ωr 2

Z " "    2 # #
R
1 2 V +v 2 V +v 1
= ρΩ abc θ0 r − r + ρbcCd0 Ω2r3 dr
0 2 Ω Ω 2

"   3  2 2 #
1 V + v R V + v R 1 R4
= ρabcΩ2 θ0 − + ρbcCd0 Ω2
2 Ω 3 Ω 2 2 4

  
1 θ0 λ 1 R4
= ρabcΩ2 − λR4
+ ρbcCd0 Ω2 (2.2.15)
2 3 2 2 4

Define torque coefficient as,

Q
CQ = Torque coefficient = (2.2.16)
ρπR2 (ΩR)2 R

Combining Eqs. (2.2.15) and (2.2.16) one has

  
1 abc θ0 λ bc 1
CQ = − λ + Cd0
2 πR 3 2 2πR 4

10
c
Copyright Peretz P. Friedmann

  
σa θ0 λ σCd0
CQ = − λ + (2.2.17)
2 3 2 8

Combining Eqs. (2.1.14) and (2.2.17)

σCd0
CQ = CT λ + (2.2.18)
8

The first term in Eq. (2.2.18) is usually called the induced torque (because

it is due to induced drag) and the second term is called the profile torque.

Using the expression for the torque, Eq. (2.2.15), it is also easy to define

the power required.

P = QΩ (2.2.19)

The power coefficient is defined as,

P QΩ
CP = 3 = = CQ (2.2.20)
ρπR2 (ΩR) ρπR2 (ΩR)3

Next it is interesting to connect the expressions which have been derived

with momentum theory, in hovering flight. Recall for this case power is given

by,

P = Tv

Therefore,
Tv T v
CPi = 3 = 2 = CT λ
ρπR2 (ΩR) ρπR2 (ΩR) ΩR

11
c
Copyright Peretz P. Friedmann

since for hover


v
λ=
ΩR

Thus,

CQi = CPi = CT λ (2.2.21)

i.e. the ideal torque coefficient, in absence of friction is identical from both,

blade element theory and momentum theory.

In hovering flight,

s r
v T CT
λ= = = (2.2.22)
ΩR 2ρAΩ2 R2 2

Combining Eqs. (2.1.14) and (2.2.22) one has


" r #
σa θ0 1 CT
CT = − (2.2.23)
2 3 2 2

from which,
r
2CT θ0 1 CT
= −
σa 3 2 2

r
1 CT θ0 2CT
= −
2 2 3 σa

 2
CT θ0 2CT θ02 4 θ0 CT 4CT2
= − = − +
8 3 σa 9 3 σa (σa)2

 
θ02 θ0 4 1 4CT2
− + CT + 2 = 0
9 3 σa 8 (σa)

12
c
Copyright Peretz P. Friedmann

 
σ 2 a2 4θ0 1 (σa)2 2
CT2 − + CT + θ =0 (2.2.24)
4 3σa 8 36 0

Equation (2.2.24) is sometimes a useful quadratic equation for CT , when the

value of collective pitch θ0 is given.

Also recall Eq. (2.2.18) and combine it with Eqn. 2.2.22, thus

3
σCd0 (CT ) 2
CQ = + √ = CQ0 + CQi (2.2.25)
8 2

where CQ0 is the profile torque coefficient.

From Eq. (2.2.21) the ideal torque coefficient is

3
(CT ) 2
CQi = √
2

Another useful quantity often used in helicopter engineering is the Rotor

Figure of Merit (F.M), which is defined as, clearly this expression is for the

case of hover
3
CQideal 1 (CT ) 2
F.M = =√ (2.2.26)
CQ 2 (CQ0 + CQi )

So the ideal rotor Figure of Merit is

3
1 (CT ) 2
√ =1
2 CQi

For an actual rotor the Figure of Merit indicates the magnitude of the

losses due to non-uniformity of flow, tip loss and profile drag for a particular

rotor. For a good rotor F.M ∼


= 0.75, where “good” implies a well designed

13
c
Copyright Peretz P. Friedmann

rotor.

In the equations used above Cd0 has appeared a number of times, a good

approximate relation for Cd0 is given by,

Cd0 = 0.0081 − 0.0216α + 0.4α2 (2.2.27)

where, α is the angle of attack measured in radians. For a reasonable angle

of attack α, this normally yields a value of Cd0 ∼


= 0.012

2.3 Vertical Climb

At high rates of vertical climb, the drag of the fuselage has to be included.

However for the relatively simple situation discussed in the class, this effect

will be neglected.

Recall that for a rotor in hover, the torque coefficient is

CQH = CQ0 + λH CT (2.3.28)

and for climb or axial flight, it is

CQ = CQ0 + λCT (2.3.29)

assuming that CQ0 does not change in an appreciable manner for vertical

flight at constant axial velocity, also from equilibrium W = T (still true).

Therefore, one can combine Eqs. (2.3.28) and (2.3.29) to get,

∆CQ = CQ − CQH = (λ − λH ) CT

14
c
Copyright Peretz P. Friedmann

r
∆CQ CT
= (λ − λH ) = λ −
CT 2

Therefore,
r
∆CQ CT
λ= + (2.3.30)
CT 2

Also recall that by definition,

V +v
λ=
ΩR

V v
=λ− (2.3.31)
ΩR ΩR

For the case of axial flight one can go through momentum theory in a

manner similar to our initial derivation for the case of hover.

Using the same assumptions as was used for momentum theory for hover,

Mass flow through rotor = ρA (V + v) = ṁ (2.3.32)

Thrust from momentum equation is given by considering stations 0 − 3

T = ρA (V + v) w

Applying Bernoulli’s equation between station 0 − 1 and 2 − 3 yields again,

2v = w

Thus,

T = 2ρA (V + v) v

15
c
Copyright Peretz P. Friedmann

Figure 2.3: Geometry for momentum theory in axial flight

Therefore,
 
T V +v v
CT = 2 = 2
ρA (ΩR) ΩR ΩR

CT v
= (2.3.33)
2λ ΩR

Combining Eqs. (2.3.30), (2.3.31) and (2.3.33) yields


r !
CT
V CT ∆CQ CT 2
=λ− = + − q 
ΩR 2λ CT 2 ∆CQ CT
CT
+ 2

 2 q
∆CQ 2∆CQ CT
CT
+ CT
+ C2T −
2
CT
2
=  q 
∆CQ CT
CT
+ 2

16
c
Copyright Peretz P. Friedmann

 q 
∆CQ ∆CQ
CT CT
+ 2 C2T
∼ ∆CQ
=  q  =2 (2.3.34)
∆CQ CT
CT
+ C2T

Therefore,

V ∼ ∆CQ
=2
ΩR CT

∆CQ
Equation (2.3.34) is based on the assumption that is small compared
r CT
CT
to , which implies a low rate of climb R/C of less than V < 10 ft/sec
2
or V < 600 ft/min.

For a given R/C (desired) and a helicopter with a specific weight (CT = CW ),

the required differential torque coefficient is provided by Eq. (2.3.34).

Typical power losses for hovering rotors (single rotor):

Mechanical Gear friction 3%

Tail rotor 7%

Aerodynamic Interference (download on fuselage) 2%

Tip losses 3%

Slipstream rotation 0.2%

nonuniform inflow (no twist 4%, or taper 3%) 7%

17
c
Copyright Peretz P. Friedmann

Figure 2.4: Geometry illustrating tip losses

2.4 Rotor Tip Losses

BR is determined from balance of the two areas shown. For lightly loaded

rotors, Betz came up with a simple relation,


2CT
B =1−
b

where, b is the number of blades.

Typically,

B = 0.97
Z BR
T = dT
AR
Z R Z BR
Q= dQ0 + b dQi
0 AR

Note that only induced torque needs to be corrected. Also the correction in

the root region is not very important because the moment arm is very small.

18
c
Copyright Peretz P. Friedmann

2.5 Extended Blade Element Theory

Consider an actuator disc, in axial flight, and select an annular element

of the disc as shown in Fig. 2.5. the elemental thrust of the b blade elements

contained in the annular ring based on blade element theory is given by,
  
1 2 V + v (r)
dT = ρabc (r) (Ωr) θ (r) − dr (2.5.35)
2 Ωr

Figure 2.5: Geometry for extended blade element theory

For the same annular ring shown, the elemental thrust based on mo-

mentum theory, see Fig. 2.6, is given by,

dT = ρ2πrdr [V + v (r)] 2v (r) (2.5.36)

Equating these two different representations of elemental thrust yields


  
1 2 V + v (r)
ρabc (r) (Ωr) θ (r) − dr = ρ2πrdr [V + v (r)] 2v (r)
2 Ωr
(2.5.37)

19
c
Copyright Peretz P. Friedmann

which is a quadratic equation for v (r).


r
Rewriting the equation in nondimensional form by introducing x =
R
bc (x)
and using the definition of solidity as σ (x) = ,
πR

  
2 V + v (x)
abc (x) Ω Rx θ (x) − = 4π [V + v (x)] 2v (x)
ΩxR

Figure 2.6: Momentum theory in axial flight

  
bc (x) 2 2 V + v (x)
a Ω R x θ (x) − = 4 [V + v (x)] 2v (x)
πR ΩxR
 
2 2 V v (x)
aσ (x) Ω R x θ (x) − − aσ (x) Ω2 R2 x = 8V v (x) + 8v 2 (x)
ΩxR ΩxR
 
2 2 2 V
8v (x) + v (x) [8V + aσ (x) ΩR] − aσ (x) Ω R x θ (x) − =0
ΩxR

20
c
Copyright Peretz P. Friedmann

   
2 1 1 2 2 V
v (x) + v (x) V + aσ (x) ΩR − aσ (x) Ω R x θ (x) − =0
8 8 ΩxR

Solution of this quadratic equation yields,

  s !
V 1 2 (θ (x) xΩR − V )
v (x) = + aσ (x) ΩR −1 + 1 + 4V 2 1
2 16 σ(x)aΩR
+ V + 16 aσ (x) ΩR
(2.5.38)

Equation (2.5.38) is an important and useful equation for determining

the inflow in hover or axial flight. Once the induced velocity is known, the

inflow angle at the blade element can determined from

V +v
φ= (2.5.39)
ΩRx

Equation (2.5.38) is a completely general expression which allows one to

determine the inflow velocity for any blade planform and pitch distribution.

A number of special forms of this equation are quite useful.

For example when c = constant and the blade twist is inversely propor-
θt
tional to x, θ (x) = , where θt is blade twist at the tip, one obtains v (x)
x
constant over the disc.

  s !
1 2 (θt ΩR − V )
v (x) = V + aσΩR −1 + 1 + 4V 2 1
(2.5.40)
16 σaΩR
+ V + 16 aσΩR

Another useful relation is obtained for the case of hover V = 0 and


3
constant chord c (x) = c. Assuming that the inflow velocity at R is repre-
4
sentative of a uniformly distributed inflow for θ = constant, one has

21
c
Copyright Peretz P. Friedmann

s !
1 2θ 3 ΩR
v (x) = aσΩR −1 + 1+ 1 4
16 16
aσΩR

r !
1 24θ
v (x) = aσΩR 1+ −1 (2.5.41)
16 aσ

r !
v (x) aσ 24θ
λ= = 1+ −1 (2.5.42)
ΩR 16 aσ

which is an approximate relation for uniform inflow frequently used in aero-

elastic calculations.

2.5.1 Optimum Rotor for Hover

Here we are interested in the optimum rotor for hover including real fluid
 
effects. We are seeking αopt for max(L/D) α(Cl /Cd )max , with friction

Clopt
αopt = for all r
a

We still want v = constant over the disc. Returning to Eq. (2.5.37) for V = 0

one has,
 
1 2 v (r)
abc (r) Ω r θ (r) − = 4πv 2 (r) (2.5.43)
2 Ωr
| {z }
α(r)

we want α (r) = opt., and v = constant.

v (r)
Let θ (r) = αopt + (2.5.44)
Ωr

Therefore abcΩ2rαopt = 8πv 2 (r)

22
c
Copyright Peretz P. Friedmann

 
R
where v (r) = constant if c (r) = ct , i.e. tapered blade.
r

R
abctΩ2 Rαopt = 8πv 2
R

 v 2 abc α
t opt
=
ΩR 8πR

q
v abct αopt
ΩR
= 8πR
(2.5.45)

and combining Eqs. (2.5.44) and (2.5.45)


r
Cl 1 abctαopt
θ (x) = opt + (2.5.46)
a x 8πR

2.5.2 Rotors Hovering in Ground Effect

Consider a rotor hovering near the ground. Recall that when the rotor hovers

far from the ground, one can obtain the inflow from momentum theory, shown

in Fig. 2.7.

Figure 2.7: Momentum theory for rotor out of ground effect

23
c
Copyright Peretz P. Friedmann

When the rotor is near the ground one expects the inflow v 0 < v for an equal

amount of thrust, as shown in Fig. 2.8. One can develop a fairly simple

analytical model for this case by using an analytic image effect, which is

schematically shown in Fig. 2.9.

Figure 2.8: Rotor inflow v 0 in ground effect

The presence of ground reduces the size of the induced drag and therefore

the power requirement is lower. At a constant power setting, descent tends

to increase thrust, as shown in Fig. 2.10.

T = K1 T ∞

Out of ground effect, the torque coefficient is given by,

3
σCd0 (CT∞ ) 2
CQ = + √
8 2

In ground effect,

CT = K1 CT∞

24
c
Copyright Peretz P. Friedmann

Figure 2.9: Geometry for rotor operating in ground effect

and at constant power,

3
σCd0 (CT ) 2
CQ = + √ K2
8 2

25
c
Copyright Peretz P. Friedmann

Figure 2.10: Thrust changes in ground effect

For constant power, one can equate torque coefficients


3 3
σCd0 (CT∞ ) 2 σCd0 (CT ) 2
+ √ = + √ K2
8 2 8 2

  32   32
CT∞ 1
K2 = =
CT K1

and thus
  32
σCd0 1 CT
CQ = +√
8 2 K1

For forward flight, the effect of V , wind or forward flight velocity has a

beneficial effect, as shown in Fig. 2.11.

26
c
Copyright Peretz P. Friedmann

Figure 2.11: Effect of forward flight velocity on thrust change in ground effect

2.5.3 Descent in Axial Flight

Condition I: Power-on Vertical Climb

Figure 2.12: Slipstream while descending in axial flight

There is an established slipstream and momentum theory applies.

T = 2ρA (V + v) v

27
c
Copyright Peretz P. Friedmann

Figure 2.13: Slipstream in hover

Condition II: Hover

Again there is an established slipstream and momentum theory applies.

T = 2ρAv 2

Condition III: Partial Power Descent (Vortex Ring State)

A ring vorticity stays with the disk, possible at an angle of descent or

wind. Momentum theory can not be applied since there is no established

slipstream. The flow is not defined both from geometry and stability point

of view; this situation is shown in Fig. 2.14.

Condition IV: Turbulent Wake State

Minimum power-off rate of descent. Disk approaches a state correspond-

ing to an almost solid disk. Redefine positive V from below and v as opposed

28
c
Copyright Peretz P. Friedmann

Figure 2.14: Vortex ring state

to it in direction and almost equal in magnitude. There is no established slip-

stream and hence momentum theory is invalid. This is a flow state for ideal

autorotation. This situation is depicted in Fig. 2.15.

Condition V: Windmill Brake State

Here there is an established slipstream and momentum theory applies; This

situation is depicted in Fig. 2.16.

w = 2v

T = 2ρA (V − v) v

Applying momentum theory

ρ ρ
p0 + V 2 = p1 + (V − v)2
2 2

29
c
Copyright Peretz P. Friedmann

Figure 2.15: Turbulent wake state

Figure 2.16: Slipstream while in windmill brake state

ρ ρ
p2 + (V − v)2 = p0 + (V − w)2
2 2
ρ 2  ρ 
p1 − p2 = V − (V − w)2 = 2V w − w2
2 2
ρ 
T = ρA (V − v) w = A 2V w − w2
2

30
c
Copyright Peretz P. Friedmann

thus,

w = 2v (2.5.47)

T = 2ρA (V − v) v (2.5.48)

T 1
V = +v (2.5.49)
2ρA v

Equation (2.5.49) represents the rate of descent in the windmill brake

state, from it the minimum rate of descent can be also obtained.

dV T
=0⇒− +1 =0
dv 2ρAv 2
s
T
vmin =
2ρA
s s s
T T T 2T
Vmin = q + =2 = (2.5.50)
T
2ρA 2ρA 2ρA 2ρA ρA
 
lb-sec2
At sea level using the appropriate value of ρ0 = ρ ρ0 = 0.0023769
ft4
r r r r
2 T T ∼ T ft
Vmin = = 29.0074 = 29
0.0023769 A A A sec

This minimum rate of descent almost represents the inverse of hover. One

can also relate this to the inflow velocity of hover vH ,

s
T
vH =
2ρA

2 T
vH =
2ρA

31
c
Copyright Peretz P. Friedmann

Figure 2.17: Figure showing momentum theory in vertical descent

For vertical climb (assuming equal weight or thrust)

T = 2ρA (V + v) v

or
T
= (V + v) v
2ρA

thus

2
vH = (V + v) v

or
 
V v v
+ =1 (2.5.51)
vH vH vH

Similarly for vertical descent, see Fig. 2.17, from windmill brake state Eq.

(2.5.48) above

2
vH = (V − v) v

32
c
Copyright Peretz P. Friedmann

or
 
V v v
− =1 (2.5.52)
vH vH vH

All the results can be plotted conveniently on one curve as shown in

Fig. 2.18, where the portion of the curve denoted by “empirical” should be

assumed to be obtainable from either flight test or wind tunnel test.

Figure 2.18: Various flight regions in axial flight

In turbulent wake state,

s
Vmin T.W ∼ T
= 1.72 or Vmin T.W ∼
= 1.72vH = 1.72
vH 2ρA

33
c
Copyright Peretz P. Friedmann

Figure 2.19: Blade element theory in descending flight

2.6 Blade Element Theory of Descent Problem

Again assume small angles and c, θ0, v =constant then,


 
∼ 1 2 2 V −v
dT = dL = ρacΩ r θ0 + dr (2.6.53)
2 Ωr

and we have to redefine the inflow ratio for vertical descent as

V −v
λ= (2.6.54)
ΩR
Z R  
1 2 3 θ0 λ
T =b dT = ρabcΩ R + (2.6.55)
0 2 3 2

Thus the only difference between the vertical climb and descent is the

sign in the square brackets in Eq. (2.6.55), the positive sign convention for

this case is

↑V+ ↓V−

and
 
1 θ0 λ
CT = aσ + (2.6.56)
2 3 2

34
c
Copyright Peretz P. Friedmann

Figure 2.20: Inplane sectional forces in descending flight

From this figure it is clear that the inplane component of lift opposes the

profile drag term

dQ = r (dD0 − φdL)
Z R
Q=b (dD0 − φ(r)dL) rdr
0

V −v
φ(r) =
Ωr
Z R
1 1
Q=b { ρcΩ2 r2 Cd0 − ρacΩ2 r2 [θ0 + φ(r)] φ(r)}rdr
0 2 2
Z R  
1 2 4 1 2 2 V −v V −v
= ρbcΩ R Cd0 − ρabcΩ r θ0 + rdr
8 0 2 Ωr Ωr
 
1 1  θ0 1 V − v  V − v
= ρbcΩ2 R4 Cd0 − ρabcΩ2R4  + 
8 2 3 2 | ΩR
{z } ΩR
λ

Cd0 σ
CQ = − λCT
8
 
Cd0 σ 1 θ0 λ
CQ = − aσ + λ (2.6.57)
8 2 3 2

35
c
Copyright Peretz P. Friedmann

The only change is the sign after the first term on the R.H.S of the above

equation.

2.6.1 Partial Powered Vertical Descent

Solve torque expression, Eq. (2.6.57), for λ

 
Cd0 σ 1
λ= − CQ (2.6.58)
8 CT

For a given throttle setting CQ is prescribed

 
V −v V v vH
λ= ≡ −
ΩR vH vH ΩR

and recall that

r
vH CT
=
ΩR 2

from momentum theory

 r  
V v CT Cd0 σ 1
− = − CQ
vH vH 2 8 CT

and
    √
V v Cd0 σ 2
− = − CQ 3 (2.6.59)
vH vH 8 (CT ) 2

One can also replot the empirical portion of the curve shown in Fig. (2.18)

as done on the next page.

36
c
Copyright Peretz P. Friedmann

1.0 0.25 -1.0 -2.0

V v
vH vH

1.0

1.72
2.0

V
vH
WB TW VR

Figure 2.21: Partial plot of Fig. 2.18

In the windmill brake range, where momentum theory is valid one can

assume approximately

 
V v v
− = 1.0
vH vH vH

On the other hand, using Fig. 2.21

 
V v
− = 0.25
vH vH

for minimum rate of descent, power off and turbulent wake state,

V
= 1.72
vH

37
c
Copyright Peretz P. Friedmann

Therefore, at sea level

s r
T T
V = 1.72vH = 1.72 = 25 (2.6.60)
2ρA A

For a parachute CD ∼
= 1.4.

therefore
1
T =D∼
= ρV 2 CD A
2

s r s s r
2T 4 T T T
V = = = 1.69 = 24.5 (2.6.61)
ρACD 1.4 2ρA 2ρA A

Comparing Eqs. (2.6.60) and (2.6.61) it is evident that they are very close

to each other.

2.7 Autorotation

It is assumed that it is a power-off condition. This condition is governed

by energy balance between losses due to friction and kinetic energy due to

rotation. Power or moment coefficient given by

Cd0 σ
CQ = − λCT = 0
8
Cd0 σ
= λCT
8

From blade element theory in descending flight


 
1 θ0 λ σCd0
CT = aσ + = (2.7.62)
2 3 2 8λ

38
c
Copyright Peretz P. Friedmann

this yields a quadratic equation for the inflow λ

2θ0 λ Cd0
λ2 +
− =0
s3  2a
2
θ0 θ0 Cd
λ=− + + 0 (2.7.63)
3 3 2a

Once λ has been determined, assuming the pitch setting is known, CT can

be determined from blade element theory, Eq. (2.7.62), by definition

T
CT =
ρA(ΩR)2
T
T = W and for a rotor A is known, this is known and thus (ΩR) can be
A
obtained.

Recall that the definition of λ for descending flight is


   r
V −v V v vH V v CT
λ= = − = −
ΩR vH vH ΩR vH vH 2
 
V v λ
− =q (2.7.64)
vH vH CT
2

Using Eq. (2.7.64) one can use Fig. 2.21 to determine in which region the

rotor is operating. If one is in turbulent wake state, one needs to use empirical

part of the curve, if one is in the windmill brake state then one can use

momentum theory.

T
= (V − v) v → momentum for windmill brake state
2ρA

since ΩR is known
 
T CT V −v v 
2
= = (2.7.65)
2ρA(ΩR) 2 ΩR ΩR

39
c
Copyright Peretz P. Friedmann

From this discussion λ is known, what is not known is V, v, Eqs. (2.7.64) and

(2.7.65) provide a solution. However θ0 needs to be adjusted so that one is

in the proper region, windmill brake.

Next, consider the equilibrium of the autorotating blade under the influ-

ence of forces acting on the blade. Recall the Fig. 2.19 that depicts blade

element theory in descending flight.

Blade at any station r has two different forces acting on it.

Figure 2.22: Geometry of a blade in autorotation

1
dL ∼
= ρΩ2 r2 a [θ0 + φ(r)] cdr
2
1
dD0 = ρΩ2 r2 cCd0 dr
2
V −v
φ(r) =
Ωr

At some station re , the inplane component of profile drag balances the inplane

component of lift, i.e.

dD0 − φe dL = 0

40
c
Copyright Peretz P. Friedmann

dD0
φe =
dL

Furthermore, φ(r) is large at the root and decreases toward the tip.

1 2 2 1
ρΩ r a [θ0 + φe (r)] φe (r)cdr = ρΩ2 r2 cCd0 dr
2 2

a [θ0 + φe (r)] φe (r) = Cd0

which yields a quadratic equation for φe (r)

aφe(r)2 + aθ0 φe (r) − Cd0 = 0

Cd
φe (r)2 + θ0φe (r) − 0 = 0
a
s 
2
θ0 θ0 Cd
φe (r) = − + + 0
2 2 a
 
V −v V −v R
φe (r) = =
Ωre ΩR re

or
re λ
=
R φe (re )

this determines the location of the equilibrium point, assuming that λ is

known, from the previous consideration in this section. The effective angle

of attack is

α(r) = θ0 + φ(r)

Here, φ(r) is largest at the root, so angle of attack increases as one goes

inboard on an autorotating blade, and therefore the inboard portion of the

blade, for a region shown in Fig. 2.22 will be stalled. If stall exceeds 40%

41
c
Copyright Peretz P. Friedmann

of the blade span, autorotation cannot be sustained and the helicopter will

crash.

In practical (actual conditions) autorotation in vertical flight will usually

take place in the turbulent wake state (TW) where the flow is not smooth

and leads to a certain amount of unsteadiness in the rotor. Therefore it is

desirable to autorotate in forward flight (low forward flight speed) because

the rotor will operate under smoother conditions.

Under established autorotative conditions there is energy balance. The

decrease in aircraft potential energy per unit time is equal to the power re-

quired to sustain the rotor speed. The pilot gives up altitude, at a controlled

rate, in return for energy to turn the rotor and produce thrust.

For vertical autorotation we have seen that

dQ = (dD − φ(r)dL) rdr

This condition can exist only at two radial stations on the blade. In reality,

some stations on the rotor will absorb power from the relative airstream

and some will consume power such that the net power at the rotor shaft

is approximately zero. When assuming uniform inflow over the disk, the

induced angle of attack is,

V −v
φ(r) = tan−1
Ωr

Induced angles inboard are large, and near the tip small. Therefore the

inboard sections are producing an accelerating torque, while the outboard

42
c
Copyright Peretz P. Friedmann

stations are producing a decclerating torque; this situation is depicted in

Fig. 2.23.

Establishing stable autorotational flight requires a skilled pilot, The rotor

RPM and rate of descent are controlled by judicious selection of the collective

pitch angle θ0.

Figure 2.23: Regions of the blade in autorotation

Autorotation in forward flight is a more realistic case that in vertical

43
c
Copyright Peretz P. Friedmann

descent. The energy balance remains the same, essentially, however, due to

forward flight there is a loss of symmetry (axial symmetry) in the induced

velocity and angle of attack. This tends to move the portions of the rotor

disk that consume power and absorb power as shown in Fig. 2.24.

Figure 2.24: Autorotation in forward flight

The basic physics of the autorotational problem remain the same. The

forward flight speed is controlled by the pilot using the cyclic controls. The

flight conditions that will allow safe entry to an autorotation and recovery

of the helicopter are usually summarized in the form of height velocity of

“H-V” curves. These are often called the “dead man’s” curves for obvious

44
c
Copyright Peretz P. Friedmann

reasons. Figure 2.25 depicts a typical H-V curve for a single engine helicopter,

such curves are provided to the pilot in the flight manual. The curves that

define the “avoid” regions are established through systematic flight tests

prior to certification of the helicopter. The tests are conducted at altitude

relative to a virtual “floor” by incrementally approaching the combination

of airspeed and altitude where acceptable autorotational capability becomes

questionable, based on pilot’s rating.

Figure 2.25: Height-velocity curve for a typical single engine helicopter

The actual size and shape of the H-V curve depends on many factors,

including characteristics of the helicopter gross weight, disk loading, number

of engines and operational density altitude.

45

You might also like