You are on page 1of 145

LECTURE NOTES

ON
HELICOPTER AEROMECHANICS























Rotorcraft Aeromechanics

2
Pre-requisite:
Aerodynamics, Flight Mechanics and Engineering Mechanics
Course objectives:
- Perpetuate, cultivate and advance the understanding of a still unusual and very
capable aircraft: the helicopter
- Learning the first principles of helicopter flight
- Using the knowledge framework from the previous years to understand the
multidisciplinary aspects of helicopter (aerodynamics, structural dynamics,
performance, Aeroelasticity, and optimisation)
- To understand basic and advanced concepts related to aerodynamic loads, vehicle
performance, basic rotor dynamics, and control of helicopters and tilt-rotor aircraft
(i.e. VTOL aircraft).
- To develop the students' understanding of helicopter aerodynamics.
- To develop the students' understanding of momentum theory and blade element
theory in estimating helicopter performance.
- To develop the students' understanding of helicopter performance problems and how
to estimate the performance of an example helicopter.

Program outcome:
- Understand the characteristics of helicopter rotor flow fields for all phases of
helicopter flight, hover, vertical flight, forward flight, autorotation, etc.
- Understand theoretical and empirical techniques used to analyse and predict the
aerodynamic characteristics of helicopters in hover, vertical flight and forward flight.
- Understand helicopter main rotor and tail rotor aerodynamic design considerations
including airfoil selection and rotor configuration trade-offs.
- Understand momentum theory uses and limitations and apply it to the estimation of
rotor performance of a selected helicopter.
- Understand and apply rotor blade element theory to estimate rotor performance and
trim conditions for a selected helicopter.
- Understand and apply combined momentum and blade element methods to analyse
helicopter rotors.
- Understand and apply empirical corrections to both momentum and blade element
techniques.
- Understand and apply component drag build-up techniques to the estimation of the
total drag of a selected helicopter.
- Understand and apply performance estimation techniques to hover, vertical climb,
autorotation, range and endurance.




Rotorcraft Aeromechanics

3
References:

1. J. Gordon Leishman, Principles of Helicopter Aerodynamics, Cambridge
University Press, 2000

2. Prouty, R.W., Helicopter Performance, Stability and Control, R.E. Krieger Pub.
Co. Florida, 1990

3. Seddon, J., Basic Helicopter Aerodynamics, B.S.P. Professional Books, 1990

4. Johnson, W., Helicopter Theory, Princeton University Press, New Jersey, 1980

5. Mil, M. et al, Helicopters Calculation and Design: Vol. I Aerodynamics,
NASA TTF-494, 1967.

6. Mil, M. et al, Helicopters Calculation and Design: Vol. II Vibrations and
Dynamics, NASA TTF-519, 1968.

7. Gessow, A. and Meyers, G.C., Aerodynamics of the Helicopter, Dover
Publication

8. Bramwell, A.R.S., Helicopter Dynamics, Edward Arnold Pub., London, 1976

9. Stepniewski, W.Z. and and Keys, C.N, Rotary Wing Aerodynamics, Vol. I and
II, Dover Publications, 1984.

10. Padfield, G.D., Helicopter Flight Dynamics: The Theory and Application of
Flying Qualities and Simulation Modeling, AIAA series, 1996








Rotorcraft Aeromechanics

4
Pitch Angle ( ):
The blade pitch angle () is the angle between the plane perpendicular to the rotor shaft and
the chord line of a reference station on the blade. For a hovering helicopter, angle of
incidence (i) is different from . As the rotor blade rotates, a downward velocity (
i
v ) is
induced. The resultant velocity
R
V is a combination of this induced velocity ( )
i
v and the
linear velocity (r) in the plane of rotation at a distance r from the hub. The angle between
induced velocity ( )
i
v and the linear velocity (r) is defined as inflow angle and the angle
of incidence (i) is reduced from by the inflow angle .

Azimuth Angle ( +):

The helicopter rotor blade moves through 360 degree azimuth. The azimuth position is
measured positively in the direction of rotation from its downstream position.


Collective change of Pitch
Within limits of stall, lift coefficient increase with increase in . If the pitch of all the blades
is increased simultaneously, the overall lift and hence thrust increases. Therefore, changing
the thrust to values more than or less than weight will cause the helicopter to climb or
descend. The means of achieving this change of pitch of all blades simultaneously is called
collective pitch change.
Rotorcraft Aeromechanics

5

Cyclic change of pitch
With cyclic pitch lever, the pilot can increase the blade pitch at one azimuth position (A) and
decreases it at a diagonally opposite position (B). As a result all the blades coming to position
A steadily have increasing pitch values those receding from A and going to B have steadily
decreasing values. This causes increased angle of attack at position A and decreased angle of
attack at position B. this cyclic variation of pitch along azimuth position is called cyclic
pitch change.














Rotorcraft Aeromechanics

6
Rotor Hinges

The development of the autogyro and, later, the helicopter owes much to the introduction of
hinges about which the blades are free to move. The use of hinges was first suggested by
Renard in 1904 but the first successful practical application of hinges was due to Juan de la
Cierva in the early 1920s. There are three hinges in the so-called fully articulated rotor:

i. Flapping hinge
ii. Drag or lag hinge
iii. Feathering hinge



Flapping hinge

The flapping hinge solves the problem of rolling moment when the helicopter is in forward
flight. In hover, pitch is maintained the same throughout the azimuth position. However,
when the rotor moves forward horizontally at a velocity V, the advancing blade (at = 90
Rotorcraft Aeromechanics

7
degree) is at a velocity V + Or and the retreating blade (at =270 degree) is at V-Or. Thus, if
the pitch is same, the advancing blade gives higher lift than the retreating blade. This
production of unequal lifts on either side of the helicopter would result in undesirable rolling
moment and excessive alternating air blades on the blade. One way of correcting this is by
setting the pitch on the advancing side lower and the retreating side higher by use of some
sort of lateral control.


Flapping Hinge with offset a
As the blade advances and develops more lift, it begins to flap upward. This then introduces a
downward vertical component of velocity in relation to the blade which reduces its angle of
incidence and hence the lift of the advancing blade. As it retreats, the opposite is true, for a
downward flapping of the blade produces an increased lift. The changes in speed in
advancing and retreating blades are compensated by opposing changes in angle of incidence
(and lift) and net rolling moment about flapping hinge becomes zero.


Advancing Side (flapping up reduces angle of incidence)
Rotorcraft Aeromechanics

8

Retreating Side (flapping down increases angle of incidence)
It is to be noted that this flapping motion is caused automatically by unequal velocities only
(i.e. without any control force by pilot) and it is referred to as aerodynamic flapping.

Drag (Lag) Hinge

The next important hinge is the drag hinge. In addition to the flapping hinge, a hinge is
essential to cater for the lead-lag motion of the blade; this is the drag hinge. The blade is
hinged about a vertical axis near the center of rotation so that is free to oscillate or lead and
lag in the plane of rotation. This flexibility makes the net moment about drag hinge zero.
Both the flapping and drag hinge (in a so-called fully articulated rotor).

Rotorcraft Aeromechanics

9

Feathering Hinge

Pitch of the blades can be increased or decreased by the pilot simultaneously or differentially
(collective and cyclic pitch change) by the use of feathering hinge.

Types of rotors

Three fundamental types of rotors have been developed so far:

Rigid rotor

In these rotors, the blades are connected rigidly to the shaft. Such rotors do not have either
flapping or drag hinge. Usually, such rotors are two-bladed.

See-saw (or teetering) rotor

Rotors in which blades are rigidly interconnected to a hub but the hub is free to tilt with
respect to shaft. These rotors are two bladed. The blades are mounted as a single unit on a
see-saw or teetering hinge. No drag hinges are fitted and therefore lead-lag motion is not
permitted. However, bending moments my still be reduced by under-stinging the rotor.

The principle of see-saw rotor is similar to that fully articulated rotor (having both flapping
and drag hinges) except that blades are rigidly connected to each other. The see-saw hinge
is like the flapping hinge located on the axis of rotation and because of rigid interconnection
between two blades, when the advancing blade, flaps up, the opposite (retreating) blade flaps
down.

Fully articulated rotor

Rotors in which blades are attached to the hub by hinges, free to flap up and down also swing
back and forth (lead and lag) in the plane of rotation. Such rotors may have two, four or more
blades, such rotors usually have drag dampers which present excessive motion about the lag
hinge.

General Expression to determine the location of K
th
blade
Azimuth of K
th
blade is given by

( )
N- Number of blades



Rotorcraft Aeromechanics

10
Rotor Provides
- Lift (thrust)
- Horizontal Propulsive Force
- Forces needed for control

Hover: Helicopters Designed to be operationally efficient there.
For forward flight the rotor is tilted forward
- Blades Flap
- Vortices create vibration and noise
- Higher forward speed: Transonic flow
o Increased drag and noise (swept blades help)
o Retreating side, High AOA, Dynamic stall, Vibration
- Forward flight speeds are limited

High Speed Forward Flight Limitations
As the forward speed increases, advancing side experiences shock effects, retreating
side stalls. This limits thrust available.
Vibrations go up, because of the increased dynamic pressure, and increased harmonic
content.
Shock Noise goes up.
Fuselage drag increases, and parasite power consumption goes up as
3
V
Rotorcraft Aeromechanics

11


Power plant Limitations
Helicopters use turbo shaft engines.
Power available is the principal factor.
An adequate power plant is important for carrying out the missions.
Momentum Theory of Rotors (Actuator Disk Theory)
Developed for marine propellers by Rankine (1865), Froude (1885).
Extended to include swirl in the slipstream by Betz (1920)
This theory can predict performance in hover, and climb.
First approximation, no details of shape
Applications:
Propellers, Rotor and Ducted Fans
Assumptions:
Rotor is modeled as an actuator disk (infinitely thin disk of area A which offers no
resistance to air passing through it) which adds momentum and energy to the flow.
Flow is incompressible (compressibility corrections can be made).
Far Upstream and Far Downstream the pressure is freestream static pressure
Flow is steady, inviscid (no drag and no momentum diffusion), and irrotational.
Flow is purely one-dimensional
Thrust loading and Inflow velocity are uniform over the rotor disk (equivalent to
assuming infinite number of blades).
There is no swirl in the wake (no rotational effect)
Low Disk Loading
Rotorcraft Aeromechanics

12

Consider an actuator disk (thin circular disk with radius of rotor radius and has infinite
number of blades. It is permeable to the air flow and supports the pressure difference on the
top and bottom surface of the disk) of area A and total thrust T. It is assumed that the loading
is distributed uniformly over the disk. Let be the induced velocity at the rotor disk and w be
the wake-induced velocity infinitely far downstream. A well-defined smooth slipstream is
assumed with and w uniform over the slip-stream cross section. The rotational energy in the
wake due to the rotor torque is neglected.
Mass flow through the rotor

(1)

By conservation of mass the mass flux is constant all along the wake.
Mass conservation

() (2)


(3)


Actual Value 0.78
Thrust from Momentum equation is given by, considering stations (0) (3)
()

(4)
Rotorcraft Aeromechanics

13
Applying Bernoullis equation from station (0) (1)


(5)
Bernoullis equation cannot apply between 1-2 stations
Below the disk, between stations (2) & (3), the application of Bernoullis equation gives


(6)

)
Because the jump in pressure is assumed to be uniform across the disk, this pressure jump
must be equal to the disk loading, T/A, that is
From equation (5) & (6)


(7)
Therefore rotor disk loading is equal to dynamic slipstream pressure.

)
(8)

)
(9)

Therefore, the static pressure is reduced by ( )( ) above the rotor disk and increased
by ( )( ) below the disk.
From equation (4) & (7)
(


(10)


(11)
From equation (4) & (11)
()

(12)


(13)


Power required to hover:

( )


(14)
Rotorcraft Aeromechanics

14

(15)


Power to hover reduces when induced velocity is small and mass flow rate through the disk is
large (large rotor disk area)



(16)

where




Pressure Variation Plot in Axial Direction:









According to momentum theory, the velocity deficit in the far wake is twice the velocity
deficit at the rotor disk.
Momentum theory gives an expression for velocity deficit at the rotor disk.
It also gives an expression for maximum power produced by a rotor of specified dimensions.
Actual power produced will be lower, because momentum theory neglected many sources of
losses- viscous effects, tip losses, swirl, non-uniform flows, etc.
Preliminary Remarks
Momentum theory gives rapid, back-of-the-envelope estimates of Power. This approach is
sufficient to size a rotor (i.e. select the disk area) for a given power plant (engine), and a
given gross weight. This approach is not adequate for designing the rotor.
Rotorcraft Aeromechanics

15

Limitations of the Momentum Theory:
The analysis made by the simple momentum theory is idealised because it neglects profile
drag losses, non-uniformity of induced flow (including the energy losses due to spilling of the
air about blade tips, commonly known as tip losses) and slipstream rotation losses. Thus an
actual rotor would require more power to hover with a given load than an ideal rotor (i.e., a
rotor having zero profile drag and uniform inflow) and therefore would be less efficient.

The order of magnitude of the rotor losses not considered by simple momentum theory,
expressed as a percentage of the total power required is as follows:

Profile drag losses: 30%
Non uniform inflow: 6%
Slipstream rotation: 0.2%
Tip losses: 3%

Lastly, it does not provide any information as to how the rotor blades should be designed for
a given thrust.

Hover Power Losses

Momentum theory gives the induced power loss of an ideal rotor in hover,

. A
real rotor has the power losses as well, in particular the profile power loss due to the drag of
the blades in a viscous fluid. There is also an induced power loss due to the non-uniform
inflow of a real, non-optimum rotor design. The distribution of the power losses of the rotor
in hover is approximately as follows:

Induced Power 60%
Profile Power 30%
Non-uniform Inflow 5% to 7%
Swirl in the wake less than 1%
Tip losses 2% to 4%

The main rotor absorbs most of the helicopter power, but there are other losses as well. The
engine and transmission absorbs 4% to 5% of the total power with turbine engines. The tail
rotor absorbs about 7% to 9% of the total helicopter power, and there is an additional loss of
about 2% due to aerodynamic interference.

Disk Loading and Power Loading
T/A or DL = Disk Loading
T/P or P/L = Power Loading
At hover T = W
Induced (ideal) power:
Rotorcraft Aeromechanics

16

()



(17)
For a single rotor helicopter in hover, the rotor thrust, T, is equal to the weight of the
helicopter, W; the disk loading is sometimes written as W/A. Disk loading is measured in
pounds per square foot. The power loading is defined as T/P, which is denoted by PL, Power
loading is measured in pounds per horsepower or newton per kilowatt. Power required to
hover is given by

. This means that the ideal power loading will be inversely


proportional to the induced velocity at the disk. The ratio T/P decreases quickly with disk
loading. Therefore, vertical lift aircraft that have a low effective disk loading will require
relatively low power per unit of thrust produced (i.e. they will have high ideal power loading)
and will tend to be more efficient.

Disk loading for helicopters are usually in the range of 100 500 N/m^2 and the
corresponding inflow velocity is in the range of 6.4 14.3 m/s (at sea level density
condition).
N/m^2
Rotor 100 500
Propeller 2500
Ducted fan 2500 - 5000
Jet 50000

The higher the disk loading, the higher the induced velocity, and the higher the power. For
helicopters, disk loading is between 5 and 10 lb/ft
2
(24 to 48 kg/m
2
).Tilt-rotor vehicles tend to
have a disk loading of 20 to 40 lbf/ft
2
. They are less efficient in hover. VTOL aircraft have
very small fans, and have very high disk loading (500 lb/ft
2
).
Power Loading
For a given gross weight, that is, a high power loading with a large value of T/P is required.
Power loading is the ratio of the thrust produced to the power required to hover, that is,

Rotorcraft Aeromechanics

17





(18)

()


(19)

This quantity should be as close as possible to the ideal value for best hovering efficiency.
Because
()


()



Maximizing the power loading requires a low tip speed ().

On the basis of simple momentum theory considerations, the ratio P/T is given by

()



(20)

To maximize the power loading (that is minimize the ratio P/T) the disk loading should be
low (i.e. the disk area should be large for a given gross weight to give a low induced velocity
and the tip speed should be low).

When using the modified momentum theory, the ratio P/T is given by



This means that the best rotor efficiency (maximum power loading) is obtained when the disk
loading is minimum and the figure of merit is a maximum.

Pure helicopters have a power loading between 6 to 10 lb/HP. Tilt-rotors have lower power
loading 2 to 6 lb/HP. VTOL vehicles have the lowest power loading less than 2 lb/HP.
Rotorcraft Aeromechanics

18

Induced Inflow Ratio
- Induced Velocity (Hover)
The induced inflow velocity, , at the rotor disk can be written as

(Non-dimensional quantity)
= Angular speed; R = Rotor radius;


Thrust and Power Co-efficient
In helicopter analysis the rotor thrust co-efficient is formally defined as


Where the reference area is the rotor disk area A and the reference speed is the blade tip
speed. All velocity components are non-dimensionalized by tip speed so the inflow ratio is
related to the thrust co-efficient in hover by

()


This is based on the 1-D flow assumption made in the preceding analysis, which means that
this value of inflow is assumed to be distributed uniformly over the disk.
If Thrust coefficient goes up, Inflow ratio goes up and if inflow goes up, power goes up. As
the goes up Thrust coefficient decreases and tip speed increases beyond the critical value.
The rotor power coefficient is defined as
Rotorcraft Aeromechanics

19


So that based on momentum theory the power coefficient for the hovering rotor is

()

()

) .


Again this is calculated on the basis of uniform inflow and no viscous losses, so is called the
ideal power coefficient. The corresponding rotor shaft torque coefficient is defined as


In non-US countries an extra is used in the denominator for



Comparison on Theory and Measured Rotor Performance
The Comparison of measured and theoretical results is shown in the above figure. In terms of
coefficients it is apparent that the ideal power according to the simple momentum theory can
be written as


Notice that the momentum theory under predicts the actual power required, but the predicted
trend that

is essentially correct. These differences between the momentum theory


and experiment occur because viscous effects (i.e. non-ideal effects) have been totally
neglected so far.
Non-Ideal Effects on Rotor Performance
Rotorcraft Aeromechanics

20
In hovering flight the induced power predicted by the simple momentum theory can be
approximately described by an empirical modification to the momentum result in


Where k is called an induced power factor. This factor is derived from physical effects such
as non-uniform inflow, tip losses, wake swirl, less than wake contraction, finite number of
blades and so on.
Induced power factor =

= Profile power;

= Induced Power


Where

is the number of blades and D is the drag force per unit span at a section on the
blade at a distance y from the rotational axis. The drag force can be expressed conventionally
as

()


Where c is the blade chord. If the section profile drag coefficient,

, is assumed to be
constant (=

) and independent of Reynolds number and Mach number and the blade is not
tapered in planform (i.e. a rectangular blade), then the profile power integrates out to be


: Solidity ratio, which is the ratio of blade area to rotor disk area. Typical values of solidity
ratio for a helicopter range between 0.05 and 0.12.
Armed with these estimates of the induced and profile power losses, it is possible to
recalculate the rotor power requirements by using the modified momentum theory result that


Figure of Merit
The figure of merit is a measure of rotor hovering efficiency, defined as the ratio of the
minimum possible power required to hover to the actual power required to hover. The figure
of merit is equivalent to a static thrust efficiency and defined as the ratio of the ideal power
Rotorcraft Aeromechanics

21
required to hover to the actual power required, that is,




Ideal power is the power required to lift the weight of the helicopter (i.e. minimum power
without any loss). Ideal power is nothing but induced power.



As thrust coefficient increases FM reaches a maximum or drops off slightly. This is because
of the higher profile drag coefficients (>

) obtained at higher rotor thrust and higher blade


section AoA.

increases as increases; behaviour depends on airfoil stall characteristics.


In practice, FM values between 0.7 and 0.8 represent a good hovering performance for a
helicopter rotor.

Because a helicopter spends considerable portions of time in hover, designers attempt to
optimize the rotor for hover (FM~0.8). A rotor with a lower figure of merit (FM~0.6) is not
necessarily a bad rotor. It has simply been optimized for other conditions (e.g. high speed
forward flight).
Using the modified form of the momentum theory with the non-ideal approximation for the
power, the figure of merit can be written as




Notice that at low operating thrusts the figure of merit is small. This is because the profile
drag term in the denominator is large compared to the numerator. As the value of


increases, however, the importance of the profile power term decreases relative to the
Rotorcraft Aeromechanics

22
induced term and FM increases. This continues until the induced power dominates the profile
term and the figure of merit will begin to approach a value of 1/k. In practice, however, the
profile drag contribution decreases this value. In practice, at higher values of rotor thrust the
profile drag (and power) increases quickly as the blade begin to stall, which will again cause
a reduction in FM.
High solidity (lot of blades, wide-chord, large blade area) leads to higher Power consumption,
and lower figure of merit. Figure of merit can be improved with the use of low drag airfoils.
If the solidity ratio () value is small FM value goes up.
We know that


If we need high figure of merit, value should be small. But in the equation (A), value
goes down value goes up and reaches stall condition. So optimum value of should be
decided to compromise both the equations.


Rotorcraft Aeromechanics

23

Induced Tip Loss
The formation of trailed vortex at the tip of each blade produces a high local inflow over the
tip region and effectively reduces the lifting capability.
Tip vortex reduces lift tip loss
B: tip loss factor
BR =

=effective radius (< R)


Adding inner root cut-out


Prandtl (vortex theory) showed that when accounting for the tip loss, the effective blade
radius,

, is given by


For climb B still holds using climb velocity
Rotorcraft Aeromechanics

24



B is shown to decrease with decreasing

(less blade-to-blade interference) and also with


increasing

(spacing of vortex sheet below the rotor). In practice, values of B for helicopter
rotors are found to range from about 0.95 to 0.98, depending on the number of blades.
Gessow & Myers (1952) suggest an empirical tip-loss factor based on blade geometry alone
where


c: tip chord
Sissingh (1939) has proposed the alternative geometric expression


Where

is the root chord of the main blade and

is the blade tip ratio (ratio of the tip chord


to the root chord).
Rotorcraft Aeromechanics

25

Blade Element Theory
Two primary limitations of the momentum theory are that it provides no information as to
how the rotor blades should be designed, so as to produce a given thrust. Also, profile drag
losses are ignored. The blade element theory provides means for removing these limitations.

The blade element theory, which was put in practical form by Drzewiecki, is based on the
assumption that element of a propeller or rotor can be considered as an aerofoil segment that
follows a helical path. Lift and drag are then calculated from the resultant velocity acting on
aerofoil, each element being considered independent of the adjoining element. The thrust and
torque of the propeller or rotor are then obtained by integrating the individual contribution of
each element along radius.

It is a strip theory. The blade is divided into a number of strips, of width dr. The
aerodynamic force lift generated by that strip, and the power consumed by that strip, are
computed using 2-D airfoil aerodynamics. The contributions from all the strips from all the
blades are summed up to get total thrust, and total power.

The resultant local flow velocity at any blade element at a radial distance y from the
rotational axis has an out-of-plane component (perpendicular to the rotor) as a result of climb
and induced inflow


An in-plane component (parallel to the rotor) because of blade rotation


The resultant velocity at the blade element is, therefore




Rotorcraft Aeromechanics

26



Rotorcraft Aeromechanics

27

The relative inflow angle (or induced angle of attack) at the blade element will be


Thus, if the pitch angle at the blade element is , then the aerodynamic or effective AoA is





()

)
For low axial velocities and a radial position sufficiently outboard from the axis
of rotation
() (

)
Aerodynamic section of the rotor blades starts from 20 to 25% of the rotor.
The elemental lift and drag, associated with a segment of width dr is


For a constant chord blade, and assuming a constant drag coefficient, the profile power
coefficient can be evaluated as
Rotorcraft Aeromechanics

28


For an accurate calculation of the profile power loss, the variation of the drag coefficient with
the angle of attack and Mach number should be included. Consider a profile drag polar of the
form


By properly choosing the constants

and

the variation of drag with lift for a given


airfoil can be well represented for angle of attack below stall. This representation for drag
coefficient was used by Bailey (1941) and his numerical example

is frequently found in helicopter calculations.


The lift dL and drag dD act perpendicular and parallel to the resultant flow velocity
respectively.
The quantity c is the local blade chord.
When is small, is large and

is small

,()

,()

)-

()

,()

)-

()


These forces can be resolved perpendicular and parallel to the rotor disk plane giving


Therefore the contributions to the thrust, torque and power of the rotor are

: Number of blades comprising the rotor


For hover and axial flight independent of azimuth angle ()

( )
Rotorcraft Aeromechanics

29

( )

( )
For helicopters the following simplifying assumptions can be made:


This is valid approximation except near the blade root, but the aerodynamic forces are small
here anyway.
The induced angle is small, so that


The drag is at least one order of magnitude less than lift, so that the contribution
( ) is negligible.
Applying these simplifications to the preceding equations results in

( )

( )
Non-dimensionalization:

()

()

()


The inflow ratio can be written as

()

/
(

)()


For a rectangular blade (c = constant)
Rotorcraft Aeromechanics

30



Similarly

()

( )
(

)()

) (


For a rectangular blade the thrust coefficient is


Torque / Power coefficient:


Since


Drawbacks of Blade Element Theory
It does not handle tip losses.
Solution: Numerically integrate thrust from the cutout to BR, where B is the tip loss
factor. Integrate torque from cut-out all the way to the tip.
It assumes that the induced velocity v is uniform.
It does not account for swirl losses.
The Predicted power is sometimes empirically corrected for these losses.


Rotorcraft Aeromechanics

31
Integrated Rotor Thrust and Power
To evaluate

and

it is necessary to predict the span wise variation in the inflow as well


as the sectional aerodynamic force coefficients

.
If 2-D aerodynamics are assumed

( )

( )
Where Re and M are the local Reynolds number and Mach number
Thrust Approximations
Based on steady linearized aerodynamics, the local blade lift coefficient can be written as

of the airfoil section


For an incompressible flow

per radian

( )


For symmetric airfoil

( )



Untwisted Blades, Uniform Inflow
Zero twist


Uniform Inflow


Rotorcraft Aeromechanics

32

<

=
This can be solved iteratively to find

for a given


is almost equal to 6

gives the average angle that required during hovering condition or angle need to be
provided during hover.

is also known as blade loading or mean pitch angle required for


operation.
When the weight increases, pitch angle also increases. After certain angle, helicopter cannot
lift-off, because blade reaches stall angle.

Linearly Twisted Blades, Uniform Inflow
Many helicopter rotor blades are designed with a linear twist, so that takes the form
Rotorcraft Aeromechanics

33
()


where

is the blade twist rate per radius of the rotor.



Using this variation in (), we get

,(

]
Using at 0.75 radius (

)
()

( )

*,(

( )

- +

*,(

)- +

]
Ideal Twist
This twist distribution while not physically realizable at the root is of interest because it will
be found to give uniform inflow with the constant chord blades.


The induced velocity can be made constant across the disk by allowing the pitch angle to
Rotorcraft Aeromechanics

34
vary inversely with non-dimensional radial distance ()
So

]
Rotor Torque
Another important quantity associated with rotor behaviour is the torque, or the moment
needed to overcome the drag and keep the rotor turning at a certain RPM in steady state
conditions. One can express the elemental torque

Total torque

( )

( )


Elemental Torque due to all the blades

]

In non-dimensional form

()

)1


Rotorcraft Aeromechanics

35


If the climb velocity is constant


In non-dimensional form the climb power will be the product of velocity and the total weight
of the helicopter.
Induced velocity in climbing condition is different from induced velocity in hovering
condition.
For hovering flight

becomes


In this case at each section

is assumed to be constant. If different aerofoils used in the


rotor blade then

will become the function of span and Mach number.



[(

()


Combining the above equations
Rotorcraft Aeromechanics

36

[(

) ]

[(

) ]


The first term is usually called the induced torque (because it is due to the induced drag) and
the second term is called the profile torque.

The power coefficient is defined as

()


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

()

()


Since for hover



So


Ideal torque coefficient in absence of friction is identical from both blade element theory and
momentum theory.
In hovering flight

<

=
From which
Rotorcraft Aeromechanics

37
(

()


This is a quadratic equation for

, when the value of collective pitch

is given.
Thus


Where

= Profile torque coefficient


Ideal torque coefficient,


Another useful quantity often used in helicopter engineering is the Rotor Figure of Merit
(FM) which is defined as

)

So that the ideal rotor figure of merit is

)

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 FM
0.75 where good implies a well-designed rotor.
In the equations used above

has appeared a number of times, a good approximation for

is given by

in radians, Angle of Attack


Rotorcraft Aeromechanics

38
For a reasonable angle of attack this normally yields a value of

0.012
In hover case


k = Empirical factor which represents additional losses (+15%)
In this case we are integrating the equations from zero to one directly (i.e. along the full span
of the blade) but near the tip the above equation will not look like this. In real rotors, at the tip
there will be lift drop. But the equation gives the lift per section.

Lifting-line theory is not strictly valid near wing tips. When the chord at the tip is finite,
blade element theory gives a non-zero lift all the way out to the end of the blade. In fact,
however, the blade loading drops to zero at the tip over a finite distance because of three-
dimensional flow effects. Since the dynamic pressure is proportional to

.
Beyond 0.5 the lift value will be large. In the actual rotor near , lift drops and then to
zero value at the tip exactly. At the tip, lift becomes zero bu there will be always drag value
and non-uniform inflow. Analysis at the tip region is very complicated and it cannot be done
by Momentum Theory.
Tip-Loss Factor
Effective Blade


B = Prandtl tip-loss factor; B = 0.95 0.98


For Untwisted blades

and uniform inflow

]
For a twist

(ideal twist, = constant)


Rotorcraft Aeromechanics

39

6(

)
Because B is between 0.95 and 0.98, we find a 6-10% reduction in rotor thrust resulting from
tip-loss effects for a given blade-pitch setting under the stated assumptions.
The correct interpretation is to consider that for the same thrust the induced inflow will be
increased to a value

(or

) is increased by a factor

compared to the case with no assumed tip losses. For


untwisted blades and uniform inflow with tip losses alone, the thrust becomes

]
For ideal twist and uniform inflow, the thrust now becomes

]
Because of tip-loss effects, a real rotor will always have a higher overall average induced
velocity compared to that given by momentum theory and so the induced power will also be
increased relative to the simple momentum result.
Tip loss constitutes an additional source of non-uniform inflow, and would normally be
factored into the value of k. Using the BET, the induced power can be written as


Using untwisted rectangular blades and uniform inflow assumptions, then with a tip loss the
total power can be approximated by


Where the induced power factor k includes the effects of both tip loss and non-uniform
inflow over the remainder of the blade. In general we can write that


Rotorcraft Aeromechanics

40

Combined Blade Element-Momentum Theory
A combination of momentum theory and blade element theory has been developed later.
Identical equations may be derived by means of vortex theory, but it is believed that the
combination of momentum and blade element theory has greater physical significance and
can be easily grasped. The combined theory can be applied for performance analysis of
helicopter in hover, vertical climb and forward flight.

In the previous sections we dealt with momentum theory which gave estimates for power and
induced velocity in terms of the thrust. We also studied blade element theory, which allowed
us to incorporate features such as number of blades, airfoil section drag and lift
characteristics, taper and twist, etc. The latter, however, assumed that the flow through the
rotor disk is uniform as in the momentum theory.
Consider a small annulus segment of the rotor disk, shown.

Rotorcraft Aeromechanics

41
The annulus is at a distance y from the rotational axis, and has a width dy.

Equating differential thrust over annular area from BET and Momentum Theory. Assume that
the annular area is not affected by either the flow inside or outside that.
On the basis of simple one-dimensional momentum theory, the incremental thrust on the rotor
annulus as the product of the mass flow rate through the annulus and twice the induced
velocity at that section. In this case the mass flow rate over the annulus of the disk is
(

) (

)
so that the incremental thrust on the annulus is
(

) (

)
This has also been known as the Froude-Finsterwalder equation.

)()

()

)()

()

/ .

/ .


Therefore, the incremental thrust coefficient on the annulus can be written as

)
This assumes no swirl in the wake. Consider first the hovering state where

. The
incremental thrust and power contributions of the annulus are given by


The total thrust coefficient of the rotor is


And the corresponding induced power coefficient is


From the BET the incremental thrust produced on an annulus of the disk is

)
Equating the incremental thrust coefficients from the momentum and blade element theories
we find that
Rotorcraft Aeromechanics

42

) (

)
Which gives


This quadratic equation in has the solution
(

)

In hovering flight condition, the above equation simplifies to
()

;
The numerical implementation of combined BEM theory is identical to classical blade
element theory. The only difference is the inflow is no longer uniform. It is computed using
the formula given earlier, reproduced below:
(

)
Average Lift Coefficient
Let us assume that every section of the entire rotor is operating at an optimum lift coefficient
and the rotor is untapered.

()

()


Rotor will stall if average lift coefficient exceeds 1.2, or so. Thus, in practice, C
T
/s is limited
to 0.2 or so.
Rotorcraft Aeromechanics

43

With ideal twist the performance of the rotor can now be recalculated. If


()


A solution for the blade pitch angle can be found from
Rotorcraft Aeromechanics

44


Alternate Derivation
The elemental thrust of the

blade elements contained in the annular ring based on the


blade element theory is given by

()

[()

]
For the same annular ring shown, the elemental thrust based on momentum theory is given by
, ()-()
Equating this two expressions yields

()

[()

] , ()-()
Which is a quadratic equation for ()
Rewrite the equation in non-dimensional form

()

6()

()

7 , ()-()

()

6()

()

7 , ()-()

()

6()

()

7 , ()-()
()

6()

()

7 , ()-()

() (),

()-

[()

() () 6

()

[()

]
(

)
(


, -


The above equation is an important and useful equation for determining the inflow in hover
Rotorcraft Aeromechanics

45
or axial flight. Once the induced velocity is known, the inflow angle at the blade element can
be determined from


It 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.
CASE (A) when chord is constant and the blade twist is ideal (inversely proportional to(),
one obtains () constant over the disk
() (

)
(


CASE (B) Another useful relation is obtained for the case of hover (V=0) and constant
chord. Assuming that the inflow velocity at 0.75 R is representative of a uniformly distributed
inflow for one has
() (

)
(


() (

) :

;

()

) :

;
which is an approximate relation for uniform inflow frequently used in aero-elastic
calculations.
Optimum Rotor for Hover
Here we are interested in the optimum rotor for hover including real fluid effects. We are
seeking

for max (L/D)


Still want = constant over the disk.
Rotorcraft Aeromechanics

46

()

6()
()


() 6()
()

7
We want ()
Let
()

()


Therefore


() ()


()


Ideal Rotor vs Optimum Rotor
Ideal rotor has a non-linear twist:


This rotor will, according to the BEM theory, have a uniform inflow, and the lowest
induced power possible.
The rotor blade will have very high local pitch angles q near the root, which may
cause the rotor to stall.
Ideally twisted rotor is also hard to manufacture.
For these reasons, helicopter designers strive for optimum rotors that minimize total
power, and maximize figure of merit.
This is done by a combination of twist, and taper, and the use of low drag airfoil
sections.
Optimum Rotor
We try to minimize total power (Induced power + Profile Power) for a given T.
In other words, an optimum rotor has the maximum figure of merit.
From earlier work, figure of merit is maximized if

is maximized.
Rotorcraft Aeromechanics

47
All the sections of the rotor will operate at the angle of attack where this value of
coefficient of lift and drag are produced.
We will call this lift coefficient the optimum lift coefficient.


All radial stations will operate at an optimum value at which

is maximum
Once angle of attack is selected, we find from
.


This determines how the blade must be twisted. Variation of chord for the optimum rotor is
given by

()


Note that

is constant (the optimum value). It follows that


()








Such planforms and twist distributions are hard to manufacture, and are optimum only at one
thrust setting. Manufacturers therefore use a combination of linear twist, and linear variation
in chord (constant taper ratio) to achieve optimum performance.
Accounting for Tip Losses
We have already accounted for two sources of performance loss-non-uniform inflow, and
blade viscous drag. We can account for compressibility wave drag effects and associated
losses, during the table look-up of drag coefficient. Two more sources of loss in performance
are tip losses, and swirl. An elegant theory is available for tip losses from Prandtl.
Prandtl suggest that we multiply the sectional inflow by a function F, which goes to zero at

Rotorcraft Aeromechanics

48
the tip, and unity in the interior.

)
When there are infinite number of blades, F approaches unity, there is no tip loss.

( )


All we need to do is multiply the lift due to inflow by F (incorporation of Tip Loss model in
BEM)
Thrust generated by the annulus
( )


From BET


Resulting Inflow (Hover) will be
()

<

=
Rotors Hovering in Ground Effect
Consider a rotor hovering in close proximity to the ground. Recall that when the rotor hovers
far from the ground one can obtain the inflow from momentum theory. Because the ground
must be a streamline to the flow, the rotor slipstream tends to rapidly expand as it approaches
the surface. This alters the slipstream velocity, the induced velocity in the plane of rotor and
therefore the rotor thrust and power. Similar effects are obtained both in hover and forward
flight, but the effects are strongest in the hovering state. When the hovering rotor is operating
in ground effect, the rotor thrust is found to be increased for a given power. When the rotor is
near the ground
Rotorcraft Aeromechanics

49

In this case one can expect the inflow

for an equal amount of thrust. One can develop


a fairly simple analytical model for this case by using an analytic image effect, which is
schematically shown below


The presence of ground reduces the size of the induced drag and therefore the power
Rotorcraft Aeromechanics

50
requirement is lower. At a constant power setting descent tends to increase thrust.



Out of ground effect


In ground effect



For constant power, we can equate torque coefficients
Rotorcraft Aeromechanics

51


And thus


For forward flight the effect of V (forward flight velocity) has a beneficial effect as show in
the figure











Rotorcraft Aeromechanics

52


Axial Climb & Descent

The rotor is climbing at velocity V and therefore the flow is downward through the rotor disk.
It is assumed that the induced velocities and w at the rotor disk and in the far wake
respectively.
The mass flux is
( )
Momentum conservation gives
Thrust = Change in Momentum
( )
From energy conservation
( )
Rotorcraft Aeromechanics

53
( )

( )

)
( ) ( )
( )

)
( )
In hover


In hover, climb or descent, rotor supports the same weight. So T will be the same
( )

)
This is a simple quadratic relationship of non-dimensional inflow w.r.t inflow at hover
condition.
The solution of the above quadratic equation is

;
The positive root of the equation should be used, since the induced velocity is positive. If we
use the negative root, induced velocity becomes negative, Inflow variation becomes



Net flow at rotor disk


Rotorcraft Aeromechanics

54


From the above expressions, we can say that inflow decreases as the helicopter rise up or
climb up. Mass flow rate increases as the velocity changes.
Power Needed in Climb and Hover

It is convenient to non-dimensionalize these graphs, so that universal behavior of a variety of
rotors can be studied.
( )
<

=
Descent condition
Thrust is always up and weight is down. The rotor to support the weight it has to push the air
down. In this case is downward and V is upward. As the rotor pushes the air down, the
farfield downstream also moves downward.
The mass flux is
( )
Momentum conservation gives
Thrust = Change in Momentum
Climb Velocity, V
Power
Descent
Rotorcraft Aeromechanics

55
( )
Thrust acting on the rotor disk is upward, but on the fluid is opposite in direction.
From energy conservation
( )
( )

( )

)
( ) ( )
( )

)
( )
Momentum Theory gives incorrect Estimates of Power in Descent.
( )
<

=

In hover


In hover, climb or descent, rotor supports the same weight. So T will be the same
( )


V/v
h
(V+v)/v
h
Climb Descent
Rotorcraft Aeromechanics

56
(

)
(

)
This is a simple quadratic relationship of non-dimensional inflow w.r.t inflow at hover
condition.
The solution of the above quadratic equation is


The roots are positive in both the cases. When V is less than


becomes imaginary value. This is valid when V is more than

is positive in both the roots


When V is much larger than

the term


Gives positive value, otherwise this becomes imaginary value.
At rotor disk,


Climb (V<0)


Rotorcraft Aeromechanics

57


Descent (V>0) (V positive upward)


If we consider V<0 for descending condition
Helicopter Up (Positive) and Down (Negative)


Assume that these expressions are valid everywhere.
Inflow diagram for vertical flight is expressed in non-dimensional form. Continuous lines are
valid. In dotted line region, flow is highly mixed so it is difficult to calculate inflow value.
Dotted lines are extrapolated values. When , there is no flow through the disk.
The lines divide the plane into four regions, where the
rotor operating condition is named the normal working state (climb and hover), vortex ring
state, turbulent wake state and windmill brake state. For climb it was assumed that the air is
moving downward throughout the flow field ( ). The flow
through the disk and in the wake is downward while the flow outside the slipstream is
upward. The climb solution may be expected to be valid for small rate of descent, however,
where at least near the rotor the flow is all downward.
For the rotor in descent, it was assumed that the air is moving upward throughout the flow
Rotorcraft Aeromechanics

58
field ( ). For the upper branch of the descent solution,
however, , so the flow is downward in the far wake while it is upward everywhere
else, including outside the wake slip-stream.
In the vortex ring and turbulent wake states the flow outside the slipstream is upward while
the flow inside the far wake is nominally downward. Because such a flow state is not
possible, there is no valid momentum theory solution for the moderate rates of descent
between V = 0 and

.
The line corresponds to ideal autorotation, P=0, and is in the center of the range
where momentum theory is not valid.
At hover V=0,

becomes 1. So the curve starts from 1 in climb.

is a constant quantity
for a given rotor. It is a fixed value in climb condition. As the velocity increases the below
term decreases and reaches zero asymptotically at very high velocity value.


Expression given below is valid both in descent and climb condition


In descent, velocity is negative so the result from the above expression keeps on increases.
The climb curve will reach

technically.


Asymptotically the curve will reach the line at 45 degree. If we extrapolate to descent, the
curve will be like as shown in graph.
If


Or



Then


Rotorcraft Aeromechanics

59
becomes imaginary value. We cannot have the root less than -2.

In descent


In this case both roots are valid, because


Is smaller than


When V is less than

or

< 2


becomes imaginary value.
For each value of

, we will have two roots one above the and one below it.
The line is the ideal autorotation case here; for points the above ideal autorotation
the rotor is absorbing power and for points below it is producing power for the helicopter.
The universal inflow curve crosses the ideal autorotation line at about


(the scatter extends over roughly

to -1.8. Real autorotation occurs at a


high rate of descent, in the turbulent wake state. In the turbulent wake state the induced
velocity curve can be approximated fairly well by a straight line on the ( )


vs (

) plane. Joining the ideal autorotation intercept (

) and the
windmill brake state boundary (( )

) gives

)
(-2,-1) is the point where the root for the descent condition starts. The above equation
represents the straight line equation.
If we substitute X as -1.6 to -1.8


Rotorcraft Aeromechanics

60




We know that

()


Autorotation is the state of rotor operation with no net power requirement. The power to
produce the thrust and turn the rotor is supplied either by auxillary propulsion or by descent
of the helicopter. A component of the aircraft forward velocity directed upward through the
Rotorcraft Aeromechanics

61
rotor disk supplies the power to the rotor, in the autorotative descent of the helicopter, the
source of power is the decrease of the gravitational potential energy. More directly, the
descent velocity upward through the disk supplies the power to the rotor. The net rotor power
is zero for vertical descent in autorotation. The decrease in potential energy (TV) balances the
induced (T) and profile losses of the rotor. Neglecting the profile losses gives ideal
autorotation
( )
The velocity with which the helicopter descends will not be at ( ) . It will be
slightly at increased descent velocity. It will not be at (

) . It will be more than


this value.
Typically the descent velocity is in the range of 15 to 20 m/s in autorotation.

()


( )

()

()

()


In autorotation, shaft power = 0


If

is less or small, then autorotative descent velocity also becomes small.



The rotor in descent operates in a number of stages, depending on how fast the vertical
descent is in comparison to hover induced velocity. Four different flow states have been
Rotorcraft Aeromechanics

62
identified with descending flight.

(a) Normal thrusting state ,() () ()-
(b) Vortex ring state ,() () ()-
(c) Turbulent Wake state , () ()-
(d) Windmill brake state ,() () ()-
Assume that the induced velocity is constant and uniform over the entire disk. Velocity of the
flow inside the slipstream is same in the direction everywhere (downward). The value of
induced velocity vary depend on flight condition. This condition is defined as Normal
Working state (Flow everywhere inside and outside the slipstream is same in the direction).

Normal thrusting state:

The normal working state includes climb and hover. For climb, the velocity throughout the
flow field is downward with both V and positive. From mass conservation it follows that
the wake contracts downstream of the rotor. A wake model with a definite slipstream is valid
for this flow state and momentum theory gives a good estimate of the performance. There
will also be entrainment of air into the slipstream below the rotor and some recirculation near
the disk, particularly for hover. Although such phenomena are not included in the momentum
theory model, their effect on the induced power is secondary.
Hover (V=0) is the limit of the normal working state, still momentum theory models the flow
well in the vicinity of the rotor disk and hence gives a good performance estimate even
though hover is nominally a limiting case

Vortex ring state:
When the rotor starts to descend, a definite slipstream ceases to exist because the flows inside
and outside the slipstream in the far wake want to be in the opposite directions. Therefore,
from hover to the windmill brake state the flow has large recirculation and high turbulence.
Sometimes this entire region is called the vortex ring state.
Descent condition is split into three regions, depending on the kind of flow. Vortex ring state
begins when helicopter just started descending from the hovering condition. In descent
Rotorcraft Aeromechanics

63
condition, the velocity of the flow ( ) is smaller than the inflow. is larger than V. So it is
directed downward( ) .

V = Farfield Upstream; = Farfield downstream velocity; = Velocity at the
disk

During the descent, flow inside the slipstream will not be in one direction. It becomes mixed
flow. As the rotor descends down and the flow goes up, rotor generates vortex flow from
bottom to top. When the strength of the vortex becomes large enough, it detaches from the
rotor (Vortex Ring State).

V = Negative because flow is coming from down to up.
= Positive because downward (Flow at the rotor disk is down)
= Positive because downward

In vortex ring state condition, the flow is downward and in turbulent wake state condition, the
flow is upward. There is a condition at which the flow at the rotor disk is zero (i.e. no flow at
the disk)
We know already that, the power is split into three terms.


When

, Power (P) becomes zero, but still the rotor is supporting the weight. Power
supplied to the rotor disc to generate thrust decreases as the descent velocity increases and
becomes zero at one particular condition.
The vortex ring state is defined by P = T (V+) > 0, so that the power extracted from the air
stream is less that the induced power.


The region with P = T (V+) < 0, is called the turbulent wake state. Partial power descents
occur in the vortex ring state. Equilibrium autorotation will usually occur in the turbulent
Rotorcraft Aeromechanics

64
wake state. At small rates of descent, recirculation near the disk and unsteady, turbulent flow
above it begin to develop. The change in flow state for small rates of climb or descent is
gradual; the momentum theory solution remains valid for some way into the vortex ring state.
Turbulent Wake State
In the turbulent wake state condition, the velocity (V+) is in the downward direction and
(V+) is in the upward direction, so there is a circulation of flow inside the slipstream (wake
region). In this condition, vibration will not be as high as in condition (vortex ring state).
When the descent velocity increases more, V becomes negative. At the rotor disk, inflow is
down, is less than the velocity V (net value is upward in direction). Farfield downstream
velocity is positive in this case also.
When the rotor had no profile power losses, power-off descent would be in this condition,
since P = T (V+) = 0 for it. While nominally there is no flow through the disk, actually there
is considerable recirculation and turbulence. The flow state is similar to that of a circular
plate of the same area (no flow through the disk, a turbulent wake above it). The flow still has
a high level of turbulence, but since the velocity at the disk is upward there is much less
recirculation through the rotor. The flow pattern above the disk in the turbulent wake state is
very similar to the turbulent wake of a bluff body.

The condition P = T (V+) = 0 is in between the condition (b) and (c).






Rotorcraft Aeromechanics

65

Windmill Brake State

When the descent velocity is still faster than condition (c), flow everywhere inside the
slipstream is in the upward direction. Here (V+) is up and thrust is in downward direction
(i.e. rotor is generating power from the flow to the rotor like windmill).
( )


( )
Sign Convention:
V > 0 is climb; V < 0 is descent
P > 0 means power is consumed; P < 0 means power is extracted.
Consider now power-off descent in terms of the blade aerodynamic loading. The inflow ratio


is directed upward through the disk, so there is a forward tilt of the lift vector. For power
equilibrium at the blade section, the inflow angle must be such that there is no net in-plane
force and hence no contribution to the rotor torque.
Rotorcraft Aeromechanics

66

The rotor efficiency is about as high as possible; a low descent rate can be achieved only with
very low disk loadings.
In normal case power is supplied to the rotor to support the weight. Power supplied to the
rotor to support the weight slowly decreases with increase in the descent velocity. At a
particular condition, (V+) becomes zero and there is no power supplied to rotor to support
the weight. After the zero velocity condition, as the velocity increases rotor generates power
from the flow to the disk. In autorotation state loss of potential energy is converted into
kinetic energy (i.e. spinning of the rotor). In autorotation state, rotors act like parachute. In
condition (a) and (d) momentum theory is valid. In condition (b) and (c), there is no theory to
calculate the inflow at this state.


Rotorcraft Aeromechanics

67
Blade Element Theory Treatment of Descent Problem

Again assume small angles and

then

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

]
Thus only the difference between the vertical climb and descent is the sign in the square
bracket. The positive sign convention for this case is () ()

]

From this figure it is clear that the in-plane component of lift opposes the profile drag term.
Torque in descending flight is given by
(

)
,


Rotorcraft Aeromechanics

68
{

()-}

]
Partial Powered Vertical Descent
Solve torque expression for
(


For a given throttle setting

(fixed by putting a particular throttle setting) is prescribed




and recall


From momentum theory and therefore
(

) (


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


On the other hand


Rotorcraft Aeromechanics

69
For minimum rate of descent, power off and

( )
Therefore



For a parachute


Therefore


Autorotation
Engine shut down (power-off), here energy balance between friction losses and rotational
energy is important


And
Rotorcraft Aeromechanics

70


Knowing get

()


From
(


Or
(


Using the above equation one can use the figure () to determine in which region one is. If one
is in TW state, need to empirical part, if in wind mill brake state use momentum theory.
Two equations with two unknowns, one has to pick

so as to get the condition one wants


()-


()


Rotorcraft Aeromechanics

71

At some station


In-plane component of profile drag balances in-plane component of lift. Further in the lift is
greater and in the outboard direction the opposite is true. At equilibrium

()-

()



Which yields a quadratic equation for

()

()

()

()

This determines .

/ location of the equilibrium section.


()

()
Angle of attack increases as one goes inboard on an auto-rotating blade. In such a blade stall
occurs inboard. If stall exceeds 40% of the blade the vehicle can be destroyed.
Forces acting on an airfoil in autorotation (Blade Element)


During the power flight condition, shaft is given by the power. When the engine fails, rotor
shaft disengages from the engine.
Rotorcraft Aeromechanics

72
If

is positive, velocity of the rotor blade decreases (since the rotor is rotating opposite
to

). If

is negative, it will accelerates the rotor blade. There is a particular angle () at


which

will be zero. If the

is zero then the rotor continues to rotate with same velocity.




Near the rotor root, is nearly zero.
() at which autorotation is
possible. Rotor is powered by air means the resultant flow rotates the rotor (Power from air to
rotor).
For power equilibrium at the blade section, the inflow angle must be such that there is no net
in-plane force and hence no contribution to the rotor torque ( ) . Because
autorotation involves induced and profile torques of the entire rotor, generally only one
section will be in equilibrium itself, while the others are either producing or absorbing power.
Inflow angle is large inboard and decreases toward the tip. Then on the inboard
sections, which produce an accelerating torque (

) on the rotor and absorb power from the


air; and on the outboard sections, which produce a decelerating torque and deliver
power to the airstream. Since there is no net power to the rotor, the accelerating and
decelerating torques must balance. For a given descent rate, the rotor tip speed will adjust
itself until this equilibrium is achieved.


may be positive, negative or zero, depending on the inflow angle . Autorotation depends
on the inflow angle (D and L also contribute to it).
When

is positive, section of the rotor blade decelerates ( decreases).
When

is zero, Autorotation (because no force in the x-direction, rotor continues to rotate
with (constant)).
When

is negative, section of the rotor blade accelerates ( increases).
Rotorcraft Aeromechanics

73


In autorotation


Condition for autorotation of a particular element is


When becomes


When

is positive


When

is zero


When

is negative


Autorotation Diagram
The figure consist of the airfoil section characteristics,

, plotted against the blade


section angle of attack,, both quantities being drawn to the same scale.
Rotorcraft Aeromechanics

74

The diagram is applied to a particular section by marking off the pitch angle of the section
from the origin of the plot along the AoA axis, and then by constructing a 45-degree line
from the pitch angle. A perpendicular dropped from any point on this line to the horizontal
axis will then define the blade-element inflow angle on the horizontal axis, in as much as
the inflow angle may be obtained by subtracting the pitch angle from the blade section angle
of attack( ). Also, because of the equality of the two legs of the right triangle, the
vertical leg is also equal to .
Consider a condition of operation as represented by point (a) is one in which

.
For this condition, the resultant force on the airfoil is displaced from the axis of rotation in
such a manner as to cause the blade element to accelerate, with a consequent increase in .
As increases, decreases. The element continues to accelerate until its rotational speed
has increased to the point where the element is operating at condition (b). Autorotative
equilibrium is established at (b), because

at that condition.
(1) For a given pitch angle,, the intersection of the 45-degree line with
(a) Any point above the curve [such as point (a)] represents an accelerating condition
wherein the resultant vector falls ahead of the rotor axis.
(b) Any point of the curve [such as point (b)] represents autorotative equilibrium
wherein the resultant vector falls along the rotor axis.
(c) Any point below the curve [such as point (c)] represents a decelerating condition
wherein the resultant vector falls behind the rotor axis.
(2) The highest possible value of the pitch angle at which autorotation may exist is such
that the 45-degree line is tangent to the curve, as at point (d).
It is important to note that autorotation is a stable phenomenon as long as the pitch angle is
less than the maximum as defined by point (d). Any disturbance which slows down the rotor
increases and accelerates the rotor to autorotative equilibrium. Similarly, if a disturbance
causes the rotor to speed up, is decreased, thereby tilting the resultant vector rearward and
decelerating the rotor to equilibrium.
If changes in the inflow velocity are neglected, then varies inversely with . On the
diagram, then, the highest rotational speed corresponds to the lowest . The pitch for
maximum rotor speed is therefore the pitch defined by the intersection of a 45-degree line
through the minimum of the

curve and the horizontal axis. As the pitch is increased,


the rotational speed will decrease more and more rapidly until the highest possible pitch for
Rotorcraft Aeromechanics

75
autorotation is reached [point (d)].
Reverse Flow
At higher rotor advance ratios, a considerable amount of reverse flow will exist on the
retreating side of the rotor disk, that is, the blade sections operate with the trailing edge into
the relative wind (at the root region .is small).
(Azimuth angle) always measured from the rear X-axis.
The locus of the region where

means that


Boundary of the reverse flow region circle is given by

Beyond this boundary is large and below this boundary is small. Consider the angle
between the line that extended from the root of the blade and Y-axis is .
In the retreating side


So
( )


This is the condition for reverse flow region boundary. When ,
In non-dimensional form

( )
( )

Rotorcraft Aeromechanics

76

The reverse flow region is a circle of diameter on the retreating side of the rotor disk. For
low advance ratio the influence of reverse flow is small, since it is confined to a small area
where the dynamic pressure is low. Therefore, up to about reverse flow effects may
be neglected. At higher advance ratios, the reverse flow region occupies a large portion of the
disk and must be accounted for in calculating the aerodynamic forces on the blade.

The normal aerodynamic force neglecting stall was given for small angle as

)
The positive directions of the various quantities are as follows:

and L upward, nose


up,

downward, and

from the leading edge to the trailing edge. In the reverse flow
region the angle of attack is


Just as in normal flow. However, in reverse flow a positive angle of attack gives a negative
(downward) lift:

|(

)
Thus an expression valid in both reverse and normal flow is

|(

)



Rotorcraft Aeromechanics

77
Momentum Analysis in Forward Flight
Consider a rotor in forward flight. In this case there is an edgewise component of velocity, as
illustrated below, because the thrust vector has to be tilted so as to provide a force component
in the direction of flight.



Rotorcraft Aeromechanics

78


The mass flow rate through the actuator disk is now

Where U is the resultant velocity at the disk and is given by
(

-
Under these conditions the axisymmetric flow that was used in the derivation of the
momentum theory in axial flight is lost. However by assumptions originally introduced by
Glauert it is possible to construct a momentum theory that allows one to calculate rotor
performance in forward flight.
The application of the conservation of momentum in a direction normal to the disk gives
(


Rotorcraft Aeromechanics

79
Using conservation of energy
(

)
(


This is the same result shown previously for the axial flight cases. Therefore


Notice that for hovering flight

, so that the above equation reduces to



This confirms that the forward flight result above reduces to the hover result as required. As
the forward flight speed increases such that

, then


This is called Glauerts High speed Approximation.
In forward flight, the rotor thrust is given by


For hovering flight


Substitute


Divide by

and rearrange to get


(

) (


Rotorcraft Aeromechanics

80
Power Required (=

,(

)-
(

)(

)
(

)
Ideal power required by the rotor is given by the product of the thrust and velocity normal to
the disc. If the ideal (induced) power required to hover and produce the same thrust

. Then

) (

)
From given T,

calculate

solve


To obtain

and then calculate

from

)
Special cases of thrust constant equations


Consider the general expression for induced velocity
(

) (


The induced velocity equation is biquadratic
(


Rotorcraft Aeromechanics

81
(

) <

=



(

)(

) (


Constant Power Equations
Variation of

and T with forward velocity

for constant power. In hover the thrust that


can be developed at a given power


With the same power the rotor at an angle of attack in a freestream velocity produces a T and
induced velocity given by the equation (for hovering rotor)


Dividing by

we get
(

) (


6(

) (

7 (


Rotorcraft Aeromechanics

82
6(

) (

7 [


Then the induced velocity in forward flight can be written as


The idea of a tip speed ratio or advance ratio (), is now used. By using the velocity parallel
to the plane of the rotor, then we define


The inflow ratio is


But it is also known from the hover case that


Finally, the solution for the inflow ratio,


Special Case,
If the disk AoA is zero, an exact analytical solution for the inflow ratio can be determined.
The induced velocity ratio in forward flight is


Squaring both sides of the above equation and rearranging gives


Dividing by

gives
Rotorcraft Aeromechanics

83
(


This quadratic has the solution

<



Notice that the induced velocity decreases quite quickly with increasing forward flight
velocity. The asymptotic approximation is obtained by letting , so that in high-speed
forward flight equation


Gives

)
For most purposes, this approximation is satisfactory for

, which in practice is the


case for
Because AoA can never be zero in any practical case, inflow equation is usually solved
numerically. Analytic solution of the inflow equation is possible, by proper selection of the
correct root from a set of multiple complex roots. There are two common numerical
approaches: 1. A simple fixed-point iteration and 2. Newton-Raphson iteration. In general, it
is necessary to solve a quadratic equation for based on the equation


Rotorcraft Aeromechanics

84
The above equation must be solved numerically
1. Trial and Error Method ( n = iteration number)


The starting value for

is usually the hover value


The error estimator is



Normally, convergence is said to occur if . One normally finds
that between 10 and 15 iterations are required with fixed-point iteration approach.
An iterative procedure for calculating inflow value may be developed by considering
Newton-Raphson solution of ()
()

6
()

()
7

()

)


Starting value


Three or four iterations are usually sufficient, starting from the value

/

As increases the induced part of the total inflow decreases dominated by term at
higher advance ratios.

Rotorcraft Aeromechanics

85


Velocities and Forces on a blade element at azimuth
Horizontal Force
H-force is the force in the plane of the disk pointing to rear. It is composed of components of
forces shown in the figure.
( )
Note H force acts in no feathering plane and is positive in the aft direction.
Rotorcraft Aeromechanics

86





Rotorcraft Aeromechanics

87

Advancing Side

Retreating Side


Rotorcraft Aeromechanics

88

force is caused by the difference in inclination of the lift force on the left and right hand
side of the hub.

force is caused by the difference in profile drag on the advancing side and retreating side.
Total H-force is usually small but can be of either sign.
Note:


Rotorcraft Aeromechanics

89
Forward Flight Performance:
For a helicopter in forward flight, the total power required at the rotor, P, can be expressed by
the equation


Where

= Profile Power;

= Parasitic Power;

= Climb Power;

= Drag
Power
Induced Power:
Induced Power is the power required to lift the total weight of the helicopter.
Profile Power:
Profile power is the power required to overcome viscous losses at the rotor or drag due to
friction on blades.
Parasitic Power:
Parasitic power is the power required to overcome the drag of the helicopter.
Climb Power:
Climb Power is the power required to increase the gravitational potential of the helicopter.
Drag Power:
The power required to overcome the difference in profile drag forces between the advancing
and retreating side (

).
Profile Drag Power
Power required to overcome the profile drag on blades

( )


In simple power calculations

is frequently neglected, however it is a considerable error.


Power consumed by one blade element is


Rotorcraft Aeromechanics

90
Where


The average power is given by


The resulting expression cannot be integrated in closed form.
Integration can be performed numerically

) ( )

-
The above equation is the profile power coefficient that includes radial flow effect.
In general the equation becomes

-
Where the numerical value of K varies from K = 4.5 in hover to K = 5 at , depending
on the various assumptions and /or approximations that are made. In practice, usually average
values of K are used that are independent of advance ratio. Bennett (1940) used an average
value of K = 4.65, while Stepniewski (1973) suggests K = 4.7. Either value will be acceptable
for basic performance studies at the advance ratio typical of conventional helicopters, that is,
for . At higher advance ratios, experimental evidence suggests that profile power
grows more quickly than the given equations. This is a result of radial and reverse flow, as
well as compressibility effects of the rotor.
The profile power coefficient without radial flow effect is

-
Consider a detailed evaluation of profile power with and without the radial flow terms:


Rotorcraft Aeromechanics

91

[( )

0.

/1

/ ]

/ ]

/ ]

()

/ ]

/ ]


If the radial flow components is neglected, then

0.

/ 1

[.

,()

()

()

]
(

()

)

Rotorcraft Aeromechanics

92
Bennett Approximation

)
We have used the relations in the above derivation


At high advance ratios, a considerable amount of reverse flow will exist on the retreating side
of the rotor disk, that is, the blade section operate with the trailing edge into the relative wind.
The locus of the region where

= 0 means that

( )
Which has the solution . Therefore, the reverse flow region where


covers a circular region of the disk diameter , with the circle Centered at ( )
( ). In this reverse flow region the sign of both

and the sectional drag


contribution to the rotor drag changes and so this must be accounted for in the radial and
azimuthal integration to find the rotor power and drag coefficients.
The effect of reverse flow can be included by changing the sign of the drag in the region
defined by on the retreating side of the disk. This may be treated by writing the
integral in the profile power equation as the sum of two parts, that is

( )

( )

( )


Where the first integral is simply the result evaluated (without reverse flow effects) and the
second integral is the increment that accounts for the proper sign on the drag inside the
reverse flow region. Evaluating the second integral gives

( )


Therefore the profile power coefficient becomes,
Rotorcraft Aeromechanics

93

)
Parasitic Power
The parasitic power is a power loss as a result of viscous shear effects and flow separation
(pressure drag) on the airframe, rotor hub and so on. Because helicopter airframes are much
less aerodynamic than their fixed-wing counterparts (for the same weight), this source of drag
can be very significant.

)
Where

is some reference area and

is the drag coefficient based on this reference area.


In non-dimensional form this becomes


Where A is the rotor disk area and f (

) is known as the equivalent wetted area or


equivalent flat-plate area. This parameter accounts for the drag of the hub, fuselage, landing
gear and so on, in aggregate.
Let f be an equivalent flat plate area representative of a helicopter fuselage. Also recall that
the drag of a flat plate, perpendicular to the flow can be represented by a drag coefficient


This expression can be built up, using conventional drag coefficient

based on the
appropriate areas of the components.



Recall

)
A reasonable number for a helicopter is


Rotorcraft Aeromechanics

94
Climb Power
The climb power is equal to the time rate of increase of potential energy. If the potential
energy is denoted as E, then E = W.h. The rate of increase of potential energy is


Where W is the helicopter weight and


Where

= non-dimensional rate of climb


Induced Power
For rotor in forward flight, the induced power of the rotor

(neglecting dynamic stall and


compressibility) can be approximated as


If the forward velocity is sufficiently high say , then the induced velocity can be
approximated by the asymptotic result predicted by Glauerts High Speed flight formula.
Therefore the power equation can be written more simply as,


Where k is the empirical correction factor to account for tip losses and non-uniform inflow
The mean value of k is around 1.15 to 1.25 sufficiently accurate for preliminary prediction of
rotor power requirement.

()


High advance ratio, low thrust corresponding to cruise


Rotorcraft Aeromechanics

95

Drag Power

()

( )

()

()

)
Total Power


The performance theory above does not account for
1. Non-Uniform inflow effects
2. Swirl losses
3. Tip losses
It also uses an average drag coefficient. To account for these, the power coefficient is
empirically corrected.


Rotorcraft Aeromechanics

96








Rotorcraft Aeromechanics

97
Blade Element Theory in Non-Axial Flight
In rotorcraft, airfoil section of the blade is subjected to oncoming velocity (V) and normal
velocity due to inflow and motion of the blade up and down i.e. flapping (relative velocity
due to flap motion of the blade).
X and Y are the coordinates of hub plane.

and

are the rotating coordinate system. It


rotates with the blade. Blade can flap also. In the motion, blade will have all three motions
(lag, flap, elastic torsion and blade twist due to torsion).
Considering only flap motion in the forward flight is the simplest one. We are assuming that
the blade is having only the flapping motion (coming out of the hub plane)
Lift and drag will be calculated from aerodynamic theories with respect to

reference
frame.
To calculate the hub load we will transfer the aerodynamic loads on the blade from


reference frame to XYZ reference frame through the transformation of the coordinate system.
The first assumption which we are making in this analysis is rotor blade is centrally hinged.
In this analysis, we are considering only (pilot input). Assume that the blades are
rigid.
In steady flight the blade motion must be periodic and is therefore capable of being expressed
in a Fourier series as


Higher harmonic terms are neglected.

Rotorcraft Aeromechanics

98


[

] [



] [

]
[

] [



] [

]
During the rotor blade design all the centers are brought to one point (mass center,
aerodynamic center and shear center). By adding the leading edge mass, C.G is brought to
25% of the chord.
Aerodynamic center at subsonic level will be normally at 25%. Pitch axis of the blade and
elastic line of the blade also passes through the 25% point on the chord. These are all due to
avoid many couplings.
In rotorcraft, airfoil is moving continuously and pitch angle also changing continuously. This
is not a static problem. Every instant we are assuming the angle of attack (for a given angle of
attack, lift and drag will be considered).
So this approximation is called Quasi-static approximation (even though pitch angle is
changing continuously). This is a dynamic situation but still we are not considering as it is.
Rotorcraft Aeromechanics

99
The velocity expressions are derived from 25% of the chord or blade.
Take a point P on the reference
The position vector will be


The angular velocity of the blade about Z-axis is .

is rotating with , which is the rotor


angular velocity w.r.t XYZ (one axis rotation). But when we consider

, it is not one axis


rotation, because it has flap motion also about

. So


about Z-axis (counter clockwise positive)
about

(clockwise positive)


r is fixed and

= Position Vector


are due to motion of the vehicle.


, -

, -


Consider as small, so = 1,
Rotorcraft Aeromechanics

100


, -

, -


is small and also small. So the product of becomes zero or neglected

()

( )

5
In non-dimensional form,

) (

) (


Every section is changing and is constant over the entire disk
Assume that

is much larger than


In


So
Rotorcraft Aeromechanics

101

is small, so

becomes small
So becomes

( )(

) (

)
From

we can derive lift and drag


( )

( ) ( )
Before integrate

and

, we have to transform it to

and

coordinate system
[

] [



] [

]
[

] [



] [

]
Assume that there is no force in the

direction.
[

] [



] [

] [



] [

]
Rotorcraft Aeromechanics

102

( )

( )
The elemental thrust

) [

)]

) [

)]
Average thrust based on the assumption of uniform inflow


Thrust is computed by integrating the lift radially to get instantaneous thrust force at the hub,
then averaging the thrust force over the entire rotor disk, and multiplying the force per blade
by the number of blades.
In forward flight the blade will undergo flapping, taking into account the first harmonic


It is convenient to change the independent variable from time to blade azimuth angle by
means of . Then since

)
Assumption for Analytical Integration
C = chord = constant
= Induced Velocity (Uniform)


= Pitch angle = constant
No cut out and No tip losses

57


Rotorcraft Aeromechanics

103

57

)5

) ]


Assumption for Analytical Integration
C = chord = constant
= Induced Velocity (Uniform)


= Pitch angle = linearly twisted rotor
No cut out and No tip losses



Rotorcraft Aeromechanics

104
Angle of Attack of a Blade Element in Forward Flight
As discussed in the class,
T
U
P
U
T
U
T
U
P
U

T
U
P
U
arctan
effecive


= ~
|
|
.
|

\
|
=
Here: sin
1s
cos
1c

R
r
tw

0
+ + + =
Vsin r sin
s
Vcos r
T
U + ~ + =
cos V r
s
R cos
s
Vcos r R
P
U + + ~ + + =


Here

means inflow in the shaft plane. This is made of two components: free-stream
component Vsin

/ R plus the induced inflow component .


Let us compute the numerator of the effective angle of attack.
( ) ( ) cos V r
s
R sin
1s
cos
1c

R
r
tw

0
Vsin r
P
U
T
U | + +
|
.
|

\
|
+ + + + =


Simplify, group sin and cosine terms. Also note that
( ) sin
1c
cos
1s

d
d

dt
d
d
d
dt
d
= =
|
.
|

\
|

|
|
.
|

\
|
= =


One gets:

) (

) + (

)
(

) (


Here we make use of the identity:
c s TPP 1
| o o + =
Note that the numerator of the effective angle of attack contains u
1c
-|
1s
and u
1s
+|
1c
. These
always occur in pairs. As far as the rotor section is concerned, a desired effective angle of
attack may be achieved either by flapping or pitching. One degree of flapping is the same as
one degree of pitching, as far as the blade section is concerned.

Finally, the effective angle of attack is obtained as
( ) ( ) { }
( ) ( ) (
(

+ + +
+ + + O
=

=
v
TPP
V V
c s
V
s c
V V
c s s c
r
T
U
T
U
P
U
T
U
effective
o | | u | u u
| u | u u
u
o
cos
0
2
sin
1 1
sin cos
1 1
sin
0
sin
1 1
cos
1 1 0
1
Rotorcraft Aeromechanics

105

( )(

) (

) (

) (

)
(

) (


When flapping angle is zero

) (

) + (

)
(

) (

[
*

) (

) +
(

) (

) (

,( )

)
( )( )-

7
Notice that we have first, second and third harmonics terms in the above equation. These
fluctuations will be felt by the passenger/pilots as vibratory loads.
Rotorcraft Aeromechanics

106
Average thrust based on the assumption of uniform inflow


Thrust is computed by integrating the lift radially to get instantaneous thrust force at the hub,
then averaging the thrust force over the entire rotor disk, and multiplying the force per blade
by the number of blades.
Integrate with respect to azimuth angle first. Use the formulas such as


Result of Azimuthal Integration

7
Next Perform Radial Integration and Normalize

]
Note that we will get the hover expressions back if the advance ratio set to zero.
Torque Coefficients
The elemental torque is given by
( )

7
Rotorcraft Aeromechanics

107

8[

)]

79






















Rotorcraft Aeromechanics

108
Flapping Dynamics of a Rotor Blade
Response of a Fully Articulated rotor with centrally hinged blades to cyclic pitch
control

Consider an articulated rotor with centrally hinged blades mounted on a whirl tower (a test
stand used for testing rotors in hover). Consider now the mathematical analysis of a rigid
(aeroelastically stiff) rotating flapping blade. The equilibrium position of the blade is
determined by the balance of aerodynamic and centrifugal forces (CF). The contributions
from the gravitational forces on the blade are small relative to the other forces and can be
neglected. Furthermore, because the centrifugal forces are much larger than the aerodynamic
forces, the flapping (coning) angle is usually quite small (between 3 degree and 6 degree is
typical for a helicopter rotor).

The rotational speed about the axis is radians per second and is constant. Assume a uniform
mass per unit length of the blade, dm. The forces on the element are aerodynamic, inertial
and centrifugal. The contribution of this small element to the centrifugal force acting in a
direction parallel to the plane of rotation
(

) ()


If the blade is coned up at some angle , then the contribution of this small element to the CF
Rotorcraft Aeromechanics

109
acting in a direction perpendicular to the blade is
(

) ()


The centrifugal moment about the hinge is
(

) ()


Induced Velocity due to flap


Inertial load on blade will be


The inertial moment about the hinge is
() ()


And the aerodynamic moment about the hinge is
(

)
Notice that the negative sign indicates that the aerodynamic moment is in the opposite
direction to the centrifugal moment. The rotating blade will reach an equilibrium position
where the CF moment about the hinge is equal and opposite to the aerodynamic moment
about the hinge.
In general for given lift distribution




The important point to remember is that the coning angle increases in proportion to rotor
thrust and decreases inversely with centrifugal forces. In other words, increasing either blade
mass or rotor speed will decrease the coning angle.
In hovering flight, the solution for is a constant(

) independent of azimuth and this


angle is called the coning angle. Under forward flight condition as a result of the cyclically
(azimuthally) varying airloads, the blade flaps up and down in a periodic manner with respect
to azimuth.
The equation of motion can be derived by writing the moment equilibrium about the hinge
point O, yields
(

) (


The mass moment of inertia of the blade about the flap hinge is


Rotorcraft Aeromechanics

110


Natural frequency of the hinged blade is . Consider the aerodynamic load and use the blade
element theory

6() 4

57


Assume () given by cyclic pitch
()

6()

6
()

6
()

4()

5
Combining Equation (1) & (2)

(()

)
Rotorcraft Aeromechanics

111
Divide through by

to get

6
()

(()

)
The Lock number is defined as




For a typical helicopter rotor, the value of the Lock number varies from 4 to 10.
Assume a steady state solution in form of
()


Where , thus

* *


Substituting equations (4), (5), (6) into equation (3) yields

]
Steady state yields

]
And


These are known as the equivalence between flapping and feathering.


In resonance the response of the forced system lags by 90 degree after the input.
Rotorcraft Aeromechanics

112
If the flapping motion of a centrally hinged blade with uniform mass distribution in forward
flight is represented as

, the thrust moment (that is the


flapping moment created by the blade thrust) about the flapping hinge is constant around the
disk (i.e. independent of azimuth)
(

) (


Where , thus

* *

So the flapping equation can be written as

* *

The flapping response is given by


So that


* *

This means that

* *
(

) (




Rotorcraft Aeromechanics

113
Flapping Dynamics of Rotor Blade in Forward Flight
The equation of motion for a flapping blade of uniform mass with a centrally located hinge is



The position vector of a mass point on the blade, with the blade mass assumed to be
concentrated at the cross sectional center of mass. Assume that the offset is zero (eR=0).


Velocity of the element is


Acceleration of the element is

) (

) (

) (

)
(


Inertia force =

)
(



Rotorcraft Aeromechanics

114
Consider the moment only about the flapping hinge (neglect other moments)

( )

is due to centrifugal force (centrifugal stiffness)


Aerodynamic load moment will be

) (

6(

) (

6(

) (

7
Where

( ) And

() ()
After substituting the results for

and

and also assuming uniform inflow and linearly


twisted blade, the result for the flapping moment can be determined analytically.

()

<( )

:
*
;( )=

:
* *
;

()

<( )

:
*
;( )=


Rotorcraft Aeromechanics

115
:
* *
;


The lock number is defined as




For a typical helicopter rotor, the value of the Lock number varies from 5 to 10. Notice
however this value depends on the density of the air.
( )

+-

()

<:
*
;( )=

< :

;=


Recall

}
*
{

} {

}
:
* *
;

9 {

*
{

} {



Rotorcraft Aeromechanics

116
In the flap motion the damping term (
*
) comes from aerodynamic moment.
If the blade pitch motion is assumed to given by the standard collective and first harmonic cycle


And also assuming the solution for the blade flapping motion to be given by the first
harmonic such that


The general flapping equation of motion cannot be solved analytically in closed form for the
general case of ( ). Therefore, two options present themselves: 1. Solve the equation
numerically 2. Find a periodic solution
Then expanding and harmonically matching constant and periodic terms on both sides of the
derived flapping equation gives
:
* *

*
{

} {

};

}
Notice the similarity of the

equation with the

equation from blade element theory

]
From this we can say that the coning angle depends on how much the blade is loaded.

) (

] (

)
Notice that by setting ( ) (hovering flight) in the above equations the following results
are obtained:


This shows that there is equivalence between pitching motion and flapping motion. If the
Rotorcraft Aeromechanics

117
cyclic pitch motion is assumed to be


Then the flapping response will be

/
Therefore, because of the dynamic behaviour of the blade the flapping response lags the
blade pitch (aerodynamic) inputs by 90 degree which is a resonant condition (for a single
degree-of-freedom linear, time-invariant system excited at its natural frequency, there will
always be a phase lag of 90 degree between the input and the output) and is independent of
any damping.
In hovering flight
:
* *

*
2

3 ;
Notice the similarity of the above equation with the equation of motion of a single degree of
freedom spring-mass-damper system, that is,

Therefore by analogy the undamped natural frequency of the flapping blade about a hinge
located at the rotational axis is one oscillation (cycle) per revolution (1/rev)(

). The blade takes one revolution to complete one full cycle (i.e. up and down
motion)
In non-dimensionalized form,


Some interesting characteristics of the resulting flapping equation are as follows:
1. In forward flight, that is, when ( ), the equation has periodic coefficients. This
does not allow an analytical closed-form solution of the flapping equation.
2. The flap damping term, which is the coefficient associated with the
*
term is
{

}

Which is of aerodynamic origin (all the terms multiplied by the Lock number come from the
aerodynamics) and is usually very high. For the case of hover and for the natural frequency of
1/rev, the corresponding damping ratio is ( ). For a Lock number this means the
damping is 50% of the critical value; thus the blade flapping motion is stable and heavily
damped. But it takes very short time to reach the steady state. In forward flight when ,
Rotorcraft Aeromechanics

118
system never damp out, it keeps on oscillate, but helicopter will not reach to that value.

3. Finally notice that the flapping equation has been derived with respect to the plane
defined by the hub of the rotor and so both the flap angle () and the pitch control
angle () will generally be functions of the azimuth angle ().

Dynamics of Blade Flapping with a Hinge Offset
The blade is assumed to be rigid and hinged at a distance from the rotational axis. The forces
acting on an element of the blade are aerodynamic, inertial and centrifugal.
The aerodynamic lift force dL acts at a distance (r-eR) from the hinge.
The inertia force, acts at a distance (r-eR) from the hinge is
()

( )
The centrifugal force, acts at a distance (r-eR) from the hinge is
(


Taking the moments about the flap hinge gives the equation of motion


( )

( )

( )


The centrifugal moment,

can be written as

( )

( )( )

( )

( )


The mass moment of inertia about the flap hinge is

( )


Rotorcraft Aeromechanics

119

The centrifugal moment can be written as

( )


This means that the flapping equation becomes

(
( )

)} ( )


Dividing through by

gives

( )


Where

is the non-dimensional flap frequency in terms of the rotational speed, that is


( )


( )

( )

)
When using the assumption that . Simplifying further leads to

( )


Typically the value of e varies from 4 to 6% for an articulated rotor (although higher for
hingeless rotor or semi-rigid rotor), so that the natural frequency of the rotor is only slightly
greater than or 1/rev. This also means that the phase lag between the forcing and the rotor
flapping response must be less than 90 degree and the flapping displacements also now
depend on aerodynamic damping. In this case the flapping equation is

Rotorcraft Aeromechanics

120
Flapping Motion and Flapping Angles
Taking moment about the flapping hinge


The equation governing the behaviour of the flapping blade becomes

) (

)
Substituting and Equating harmonics gives

are perturbations of the disk from being parallel to the swash plate. The blade
responds in resonance 90 degree after the aerodynamic forcing.
Forward Flight Case:
Consider the rotor blade as shown in the figure. It consists of a hinge offset, denoted by e and
a torsional spring having spring


Which can be also represented by an appropriate fundamental frequency

, combined with
flapwise moment of inertia of the blade about the hinge

. Such a model can be used as an


approximate representation of Hingeless blades, furthermore when

, it represents an
articulated blade assuming that the flapping motion is restricted to small angles (

) one
can assume

The position vector of a mass point on the blade, with the blade mass assumed to be
concentrated at the sectional center of mass
Rotorcraft Aeromechanics

121



From the geometry
( ) ( )

( )


We know that

( )

( )


Where is the acceleration vector and

, and it is usually assumed that the speed of


rotation is a constant, i.e. . Denote by

the position vector of a mass point


in the rotating reference frame, then
Consider next moment equilibrium of the blade about its flapping hinge, taking clockwise
moment as positive


where

= Position vector from hinge line and

is the moment due to inertia loads and


Rotorcraft Aeromechanics

122

( ) ( )

( ) ( )

( )

[( )

( )]

[( )

( )]


The restoring moment of the spring is


Finally consider the moment due to the aerodynamic forces acting on the blade. Assume that
the lift per unit span on the blade is denoted by L which acts perpendicular to the blade as
shown, due to the small angle assumption the aerodynamic moment is simply given by

( )


For moment equilibrium about the hinge one has


Combining all the above equations one finally obtains
[( )

( )]

( )


In order to proceed with the analysis, certain assumptions have to be introduced. First let us
assume that the helicopter rotor is in forward flight. From the general, introductory comments
which have been made, it is clear that for forward flight a collective pitch setting

on the
blade is required, furthermore in forward flight it has been mentioned that cyclic pitch is
required to trim the pitching and rolling moments acting on the rotor, thus
()


where is the azimuth angle measured from the straight aft position as shown in the
figure below
Rotorcraft Aeromechanics

123

The advance ratio is defined as


where is the velocity of forward flight, R is the rotor radius and

is the rotor angle of


attack.
As a consequence of the pitch program represented by the pitch angle equation, the time
history of the flapping angle can be assumed to have a similar mathematical form thus
()

()

()


Next consider the aerodynamic loads, acting on a typical airfoil is

5(

)
Since we have assumed small inflow angles
4

5
Thus the equation reduces to

)
Next determine the velocity components, parallel and perpendicular to the plane of rotation
Rotorcraft Aeromechanics

124
(also known as the Hub-Plane). Recall that the flapping angle was assumed to be small.

From the geometry


Where the first term is due to forward flight and the second term is due to rotation
Similarly

( )


Where the first term is due to inflow ratio , the second term is due to the flapping motion
() and the last term of the above equation is due to the component of the advance ratio
along the blade. The inflow ratio is defined as


where is the induced velocity over the rotor disk, and perpendicular to the disk (positive
down).
Combining all the above equations and the lift per unit span of the blade is given by

*(

)(

)
( ), ( )(

)
(

) -+
Also during the manipulation, it is convenient to use this identity
( )

( ) ( )
By the equations combination one can obtain the following relation (harmonics above the
first are neglected)

( )

]
[

]
[

]
Rotorcraft Aeromechanics

125
Where

( )()


Where

( )()

( )


A simultaneous solution of these three equations yields the rotor blade flapping
displacements

in terms of the control displacements

for the flight


condition described by the flow quantities and , and the rotor geometry, mass and elastic
characteristics described by ()

.
In order to have some physical feel for typical values of the parameters used in the flapping
analysis of a typical rotor blade the following information is provided


Rotorcraft Aeromechanics

126


Furthermore few additional comments on these equations should be made. All the above
equations based on the assumption of uniform inflow over the disk, i.e

For the case of forward flight this inflow can be determined from

is the thrust coefficient which is closely related to the weight of the helicopter. The
assumption of constant inflow velocity the disk is not a very realistic one, a more reasonable
assumption would be
( )






Coning Angle

The coning angle

is the average blade flapping angle during one revolution of the blades.
The coning angle can be thought of as the angle between the blades and the TPP. The coning
angle remains the same whether the TPP is perpendicular to shaft or not. It is independent of
time or blade azimuth. The presence of a coning angle has been pointed out to be the angle
that results from the moment balance about the flapping hinge as a result of centrifugal and
aerodynamic forces. Because the centrifugal loads remain constant for a given rotor speed,
the coning angle varies with both the magnitude and distribution of lift across the blade. As
the magnitude of the inflow increases, for a given overall total rotor thrust the blade must
Parallel to TPP
Tip Path Plane (TPP)
Rotorcraft Aeromechanics

127
become more highly lift loaded toward the tips. This produces a higher aerodynamic moment
about the hinge and therefore a higher coning angle.


Coning Angle during Hovering

Because the blades of a helicopter rotor are hinged near the root with a flap hinge, the blades
are free to flap up or down. The lift forces will tend to flap the blades up, while the
centrifugal forces will tend to push the blades down. A static equilibrium is achieved in
hover, where the blades come to rest at an equilibrium coning angle, called |
0
. In forward
flight, the airloads tend to vary in a sinusoidal form, and a static equilibrium is not feasible.
Rather, the blades undergo a limit cycle oscillation of the form:
( ) ( ) ( ) ( ) ( )
( ) ( ) ( ) ( ) ... 2 cos
c 2
2 sin
s 2
cos
c 1
sin
s 1

... t 2 cos
c 2
t 2 sin
s 2
t cos
c 1
t sin
s 1

0
t
+ + + + + =
+ + + + + =

We will study the forward flight flapping motion in detail later in this course. In this section,
we study a technique for estimating the coning angle when the blade is in hover. For
simplicity, we will assume that the hinge point is at the origin, r=0.

|
0
dL
dCentrifugal
Force
r

Consider a small strip of the blade of width dr, at a distance r from the hinge point as shown
in the picture above. This strip is subjected to lift force dL and the centrifugal force dCF. It is
also subjected to a drag force dD, perpendicular to the plane of the paper. This drag force
only affects the lead-lag motion of the blade, and does not influence the flapping motion.
From blade element theory, the lift force dL = ( ) ( ) dr
l
C
2
r c
2
1
dr
l
C
2
v
2
r c
2
1
~
(

+
Moment arm = r cos|
0
~ r
Counterclockwise moment due to lift = ( ) dr
l
rC
2
r c
2
1

Rotorcraft Aeromechanics

128
Integrating over all such strips,
Total counterclockwise moment =
( )
}
=
=
R r
0 r
dr
l
rC
2
r c
2
1

Next, the centrifugal force acting on this strip =
( )
rdm
2

r
dm
2
r
=
Where dm is the mass of this strip
This force acts horizontally. The moment arm = r sin|
0
~ r |
0

Thus, clockwise moment due to centrifugal forces = dm
0

2
r
2

Integrating over all such strips, total clockwise moment =
0

2
I
R r
0 r
dm
0

2
r
2

}
=
=

Where I is the moment of inertia of the blade about the hinge point
At equilibrium coning angle position, the clockwise moments must counteract the
counterclockwise moments. Thus,
0

2
I =
( )
}
=
=
R r
0 r
dr
l
rC
2
r c
2
1

Solve for the coning angle to get:
0
= }
=
=
|
.
|

\
|
=
}
=
=
R r
0 r
R
r
d
effective

3
R
r
I
4
acR
I
R r
0 r
dr
l
C
3
cr
2
1

The quantity
I
4
acR

ratio between the aerodynamic forces and the inertial forces acting on the rotor blade. In
general has a value between 8 and 10 for articulated rotors (i.e. rotors with flapping and
lead-lag hinges). It has a value between 5 and 7 for hingeless rotors.




Rotorcraft Aeromechanics

129

Coning Angle in Forward Flight

Because the blades of a helicopter rotor are hinged near the root with a flap hinge, the blades
are free to flap up or down. The lift forces will tend to flap the blades up, while the
centrifugal forces will tend to push the blades down. A static equilibrium is achieved in
0
. In forward
flight, the airloads tend to vary in a sinusoidal form, and a static equilibrium is not feasible.
Rather, the blades undergo a limit cycle oscillation of the form:
( ) ( ) ( ) ( ) ( )
( ) ( ) ( ) ( ) ... 2 cos
2c
2 sin
2s
cos
1c
sin
1s

... 2 cos
2c
2 sin
2s
t cos
1c
t sin
1s

0
t
+ + + + + =
+ + + + + =

In this section, we will extend the derivation for the coning angle previously limited to blades
in hover, to blades in forward flight. For simplicity, we will assume that the hinge point is at
the origin, r=0.
|
0
dL
dCentrifugal
Force
r

Consider a small strip of the blade of width dr, at a distance r from the hinge point as shown
in the picture above. This strip is subjected to lift force dL, the centrifugal force dCF, and an
inertial force that resists the blade motion. It is also subjected to a drag force dD,
perpendicular to the plane of the paper. This drag force only affects the lead-lag motion of the
blade, and does not influence the flapping motion.
From blade element theory, the lift force dL = ( ) ( ) dr
l
C
2
r c
2
1
dr
l
C
2
v
2
r c
2
1
~
(

+
Moment arm = r cos
0
~ r
Counterclockwise moment due to lift = ( ) dr
l
rC
2
r c
2
1

Integrating over all such strips,
Total counterclockwise moment =
( )
}
=
=
R r
0 r
dr
l
rC
2
r c
2
1

Rotorcraft Aeromechanics

130
Next, the centrifugal force acting on this strip =
( )
rdm
2

r
dm
2
r
=
Where dm is the mass of this strip

This force acts horizontally. The moment arm = r sin
0


~ r
0

Thus, clockwise moment due to centrifugal forces = dm
0

2
r
2

Integrating over all such strips, total clockwise moment =
0

2
I
R r
0 r
dm
0

2
r
2

}
=
=

where I is the moment of inertia of the blade about the hinge point.
Finally, in forward flight, the blades tend to flap up and down and do not necessarily stay
fixed in space. Inertial forces are generated resisting the flap motion. These forces generate a
moment about the hinge.
As before, consider a strip of the blade with a mass equal to dm. If the blade has an angular
acceleration

, this element will have a linear acceleration r

. The accompanying inertial
force (opposing the blade motion) is dm r

. This force will act downwards if the blade is
flapping up with a counterclockwise angular acceleration that increases the upward directed
(clockwise) angular velocity. The accompanying clockwise moment for this element is
. dm
2
r

integrating over all such elements, we get the total clockwise moment attributable to
inertial effects as
I
R r
0 r
dm
2
r

=
}
=
=

At equilibrium coning angle position, the clockwise moments must counteract the
counterclockwise moments. Thus,
0

2
I I +

=
( )
}
=
=
R r
0 r
dr
l
rC
2
r c
2
1
(1)
Notice that the first term on the left hand side of (1) is the only new term compared to our
previous derivation, and accounts for inertial effects in forward flight.
The homogeneous part on the left-hand side looks like a spring-mass system of the form
0 kx x m = + (2)
It may be shown that the natural frequency of the spring mass system is
m
k
. In our case in
Rotorcraft Aeromechanics

131
equation (1), the spring constant k=
2
I and the mass m = I. Thus, the natural frequency of
the blade in flapping is just O. That is, the blade will have a natural tendency to flap up and
down exactly once per revolution.
The right hand side of equation (1) is called the forcing function and will contain a steady
part, a first harmonic (i.e. terms containing sinOt or cosOt, as well as second and higher
harmonic terms. The first harmonic terms will excite the system at its natural frequency of O.
To avoid very large flapping motions that may result when the blade is excited at its natural
frequency, a small amount of physical damping is added to the system. This is in addition to
the aerodynamic damping (i.e. aerodynamic resistive forces that are proportional to d|/dt)
that are inherently present in the system.
How do the blade dynamics behave when there is a forcing function component on the right
hand side of the form AsinOt, and a damping term on the left hand side of form c

? To find
out let us solve the equation:

t A I c I O = O + + sin
2
| | |

(3)
To solve this equation, we will assume a solution of form:
t C t B O + O = cos sin |
Plugging in the assumed solution into equation (3), we get:
t A t cC t cB O = O O O O sin sin cos
which has the solution B=0 and C = -A/c . Thus the blade will under a motion
( )
2
sin cos
t
| O
O
= O
O
= t
c
A
t
c
A

when subjected to a force of AsinOt.
In other words, the blade response will be proportional to the amplitude A of the resonance
force, but will lag the force by 90 degrees.

Longitudinal Flapping Angle
The coefficient

represents the amplitude of the pure cosine flapping motion. This


represents a longitudinal or fore aft tilt of the rotor tip path plane. In forward flight, the rotor
disk has a natural tendency to tilt back (longitudinally) because of the dissymmetry in lift
produced between the advancing and retreating sides of the disk. As a result of the higher
dynamic pressure on the advancing side of the disk, the blade lift is increased over that
obtained at . Therefore, as the blade rotates into the advancing side of the
disk, the excess lift causes the blade to flap upward, which decreases its lift.
Rotorcraft Aeromechanics

132

Over the front of the disk, the dynamic pressure reduces progressively and the blade reaches a
maximum displacement with

at . As the blade rotates into the retreating side


of the rotor disk, the deficiency in dynamic pressure causes the blade to flap downward. This
downward flapping motion increases the AoA at the blade element, which tends to increase
blade lift over the lift that would have been obtained without flapping motion. This upward
and downward flapping of the blade tends to reduce and increase the AoA at the blade
elements by an amount
( )


5

Rotorcraft Aeromechanics

133

The flapping motion is expressed as a Fourier series sum of simple harmonic motion:


The motion represented by



First Harmonic Cosine Flapping Motion
Figure shows that is maximum at and minimum at . Therefore

is the
longitudinal flapping angle. The longitudinal flapping angle is positive in the downward
direction at (aft tilt of the TPP).
Lateral Flapping Angle
The coefficient

represents the amplitude of the pure sine motion. This represents the
lateral or left-right tilt of the tip path plane. In addition to the natural tendency for the disk to
tilt back with a change in forward flight speed, the disk also has a tendency to tilt laterally to
right. This effect arises because of blade flapping displacement (coning). For the coned rotor,
the blade angle of attack is decreased when the blade is at and increased when
.
Rotorcraft Aeromechanics

134

The motion is represented by



First Harmonic Sine Flapping Motion
Figure shows that is maximum at and minimum at . Therefore

is the
lateral flapping angle.







Rotorcraft Aeromechanics

135
Reference Planes Used in Forward Flight Analyses

Hub Plane
Hub Plane is a plane perpendicular to the rotor shaft. Both cyclic pitch and flapping are
present. It is the best reference plane to use for Hingeless or bearingless rotors, dynamics,
Aeroelasticity and vibrations. It is also suitable for articulated rotors with a large hinge offset.
Airfoil () is measured from hub plane.







No Feathering Plane (NFP)
When using this plane there is no cyclic pitch. Good for performance calculations of
articulated rotors. It is useless for Hingeless and bearingless rotors. It is not in widespread use
in modern helicopter analysis. This plane coincides with the swash plate plane. Airfoil

(no
cyclic) measured from NFP.



Rotorcraft Aeromechanics

136




Tip Path Plane (TPP)






No flapping component. Airfoil angle (). Good for inflow calculations, unsteady airloads
on the rotor. Frequently, eventually one has to switch to the hub plane through an appropriate
transformation
Stability Plane
It is an inertial axis system, and is useful when motions are referred to inertial space.
Change in Reference Planes
When using the NFP we have derived the expression

]







From figure above


Rotorcraft Aeromechanics

137
Assume


For small angles

]


Rotorcraft Aeromechanics

138













Rotorcraft Aeromechanics

139
Trim Analysis
The purpose of a trim analysis is to determine the rotor control settings, the rotor disk
orientation and the overall helicopter orientation for the prescribed flight or test conditions. In
general the two basic trim procedures that have been used are:
(A) Propulsive Trim
In this trim procedure, which simulates actual forward flight conditions, the rotor is
maintained at a fixed value of the thrust coefficient with forward flight. Horizontal
and vertical force equilibrium is maintained. Furthermore zero pitching and rolling
moments are also enforced. A comprehensive version of this trim procedure accounts
also for the effect of the tail rotor. In this case yaw moment and side force equilibrium
are also maintained.

(B) Wind Tunnel Trim or Moment Trim
This trim procedure simulates conditions under which a rotor would be tested in the
wind tunnel; pitching and rolling moments or the rotor are maintained at zero. For
equilibrium is not required for this case because the rotor is mounted on a supporting
structure in the wind tunnel (known as the test stand) and it provides the appropriate
force reactions that are needed for equilibrium. Yaw moment equilibrium is also
usually provided by the test stand.

Propulsive Trim
This flight condition simulates straight and level flight at a fixed advance ratio. The
weight of the helicopter is a constant quantity W. The control parameters that can be
used to achieve trim are provided below:

Collective Pitch

controls the magnitude of the rotor thrust. The collective is


changed by a lever operated by the pilots left hand, with an upward pulling motion
required for an increase in thrust.

Cyclic Pitch

and

controls the pitching and rolling moment of the rotor. Lateral


cyclic pitch

produces rotor disk tilt left and right. This changes the orientation of
the rotor thrust vector, producing both a side force and a rolling moment.
Longitudinal cyclic

imparts a once-per revolution cyclic pitch change to the


blades, such that the rotor disk can be tilted fore and aft. Both lateral and longitudinal
cyclic are controlled by the pilot using a cyclic stick (similar to the conventional stick
on a fixed-wing aircraft) which is held in the pilots right hand.

Yaw- is controlled by using tail rotor thrust. The pilot has a set of pedals, which are
operated by the pilots feet, just like the rudder of a fixed wing aircraft. By pushing
the pedals in the required direction, the collective pitch on the tail rotor is changed,
producing a change in the tail rotor thrust, and inducing yaw to the right or left. The
tail rotor has no cyclic pitch control and therefore it operates essentially in an
untrimmed mode.

Trim is a complicated and to some extent non-linear problem, particularly at high
advance ratios. There are also many cross coupling effect which are unavoidable and
usually cross coupling effects are minimized by the appropriate design of the
electronics (automatic pilot, control-feedback and mechanical or structural design of
Rotorcraft Aeromechanics

140
the rotor).

The simplifying assumptions made are listed.

Use blade element theory for the calculations of the aerodynamic loads, with certain
assumptions for the wake inflow to calculate the blade flapping and control angles.

The principal approximation here is the neglect of blade flexibility and unsteady
aerodynamics.

Neglect reverse flow and dynamic stall, these effect are neglected here for
convenience (from a calculation-point of view) and should not be neglected when
dealing with an actual problem.
It is rare that a trim solution can be obtained in closed form; usually an iterative/numerical
procedure is used to obtain trim.
Mathematically when the propulsive trim problem is considered the known (or given)
quantities are:
W Weight of the Helicopter
= Advance ratio in straight and level flight
The quantities that have to be obtained in the process of solution are:


The geometry of the problem is illustrated in the figures:
Figure 1 shows the forces and moment acting on the helicopter in the longitudinal plane and
Figure 2 shows these quantities for the lateral plane. It is convenient to decompose each
contribution (force or moment) according to its source. Thus the various contributions are
identified by subscripts, such as:
MR Main Rotor; F Fuselage; HT Horizontal Tail; VT Vertical Tail; TR Tail Rotor;
and other sources are identified by O. For example the pitching moment can be written as


In the description of the trim procedure that follows the Hub Plane (HP) is used as the
reference plane. The flight path (FP) angle is

. It will be also assumed that there is no


sideslip, thus the fuselage side force

contribution can be assumed to be negligible. Also, it


will be assumed for simplicity that there are no contributions from the horizontal and vertical
tails
Rotorcraft Aeromechanics

141

Vertical force equilibrium yields:


Longitudinal force equilibrium yields:


Lateral Force equilibrium yields:


Rotorcraft Aeromechanics

142


Pitching Moment equilibrium about the hub yields:

) (

)
Rolling Moment equilibrium about the hub yields:

)
Torque equilibrium about the shaft yields:


Using small angle assumptions allows one to simplify the above equations; i.e. all angles
used in the above equations are small


The last two terms are negligible compared to the first two, thus


Carrying out similar operations on the other equations yields

) (

)
Assuming

allows one to simplify the last equation


Recall that the thrust is simplify the average of the blade lift during one revolution, multiplied
by the number of blades


The thrust coefficient can be written as

()

6(

) (

)7


Because of the complexity of

this equation must be usually treated numerically.


Rotorcraft Aeromechanics

143
However, by assuming uniform inflow, that is ( ) over the disk, and c-
chord constant;

; and linear twist given by

]
The rotor torque, side force, drag force and moments about the appropriate axes can be
computed by a similar manner.
The rotor drag force also denoted often as the drag force


Or in non-dimensional form

6(

) (

) (

7 6(

) (

)7


The rotor side force also known as the Y-force


The rotor torque is given by


The rotor rolling moment is given by


The rotor pitching moment is given by


The above equations can be manipulated further. To complete the formulation of the trim
problem one need appropriate relations for the inflow on the main rotor and the tail rotor.
Assuming uniform inflow distribution


Rotorcraft Aeromechanics

144
It is common practise to set


However, for the tail rotor


When the side-slip angle is zero

and


Solution of the propulsive trim problem implies solving all the above equations together with


Some additional comments on the solution of the Trim equations
When solving the trim equations the quantities

are provided. Thus


( )



Similarly

( )






Rotorcraft Aeromechanics

145

You might also like