You are on page 1of 56

LECTURE NOTES

ON
HELICOPTER AERODYNAMICS

Academic Program On
Aircraft Engineering, Avionics and Manufacturing Technology
for HAL Design Trainees

G. Bandyopadhyay
Professor

Department of Aerospace Engineering


IIT Kharagpur
INTRODUCTION

Landmarks in the historical development of helicopter

15th Century: Leonardo da Vinci sketched a machine for vertical flight using a screw type
propeller.

18th Century: Sir George Cayley constructed models powered by elastic elements and made
some sketches.

19th Century: The problem of cheap, reliable and light engine is still not resolved. W. H.
Philips (England, 1842) constructed a 10kg steam powered model.

Enrico Forlanini (Itali, 1878) built a steam driven model. It climbed to a height
of 12m and stayed aloft for about 20 minutes.

20th Century: (i) Renard (France, 1904) built a helicopter using a two cylinder engine.

(ii) Paul Cornu (France, 1907) constructed the first man-carrying helicopter
with two contra-rotating rotors of 6m diameter. The weight was 260kg.
It used a 24Hp engine. It achieved a height of 0.3m for about 20seconds.
First successful manned helicopter flight.

(iii) Juan de la Cierva (Spain, 1920-30) developed autogyro. A propeller is


used for propulsive force. A rotor is developed for generating lift.
Cierva incorporated flapping hinges in his design. He was the first to
use flapping hinges successfully.

(iv) Igor Sikorsky (USA, 1939-41) built the modern helicopter (VS-300) in
1941. It was a helicopter with a 3-bladed main rotor and a tail rotor.
Several hundreds were produced.

(v) Piercy (1977) developed the shock-free transonic aerofoil and it became
possible to increase the forward speed of helicopter greatly.

2
Leonardo da Vinci’s vertical-lift machine, 15th century.
Courtesy NACA

Sir George Cayley’s helicopter and airplane, 1976.


Courtesy NACA.

3
Genealogical Tree

Flying machines can be broadly classified, as given below,

Flying Machines

Lighter than air Heavier than air

Air ships Fee Balloons Kit Balloons

Fixed Wing Aircraft Rotary wing Aircraft

Glider Sea-plane Land plane Amphibian

Autogyro Helicopter

Differences between Autogyro and Helicopter

Autogyro Helicopter

1. Propeller connected to engine provides 1. Engine driven motor provides both


thrust. Rotor driven by airflow thrust and lift
provides lift 2. Can take-off vertically
2. Can not take off vertically 3. Can hover
3. Can not hover 4. Can only be done by suitable blade
4. In case of engine failure parachute pitch change.
effect is achieved without blade pitch
control and it can glide safely

4
Thrust Lift Thrust Lift

Autogyro Helicopter

Use of helicopters

Helicopters are used for both military and civil purposes :

Military : 1. Defense helicopter for direct support of infantry. These


Helicopters are provided with armaments such as machine
guns cannons, missies etc.
2. Air observation post
3. List Reconnaissance and communications
4. Search and Rescue (where hover is essential)
5. Anti-submarine activity

Civil : 1. Used as crane for constructional work of structural assemblies


2. Patrolling of highways, oil pipelines, electrical transmission lines
3. Forest patrolling, forest fire extinguishing
4. Agricultural operations in planting
5. Emergency rescue and medical aid.

Comparison between Helicopter and Aircraft

Helicopter Aircraft
1. Helicopter is less efficient in power and fuel 1. It is more efficient in power and fuel
requirements. requirements.

2. Maximum speed achieved is 400 km/hour. 2. Speed achieved can be very high.

3. It can hover. 3. It can not take-off vertically

4. It can take off vertically. 4. It can not hover. In fact, a minimum speed
(stalling speed) is required.

5
Helicopter configurations
Helicopters can be classified based on rotors as shown below:

(1) Single rotor helicopter (with tail rotor) (2) Side-by-side helicopter

(3) Coaxial contra-rotating (4) Tandem Overlapping

(5) Tandem

6
Chapter - 1

Characteristics of Main Rotor

Characteristics of the main rotor in „Single–Rotor‟ helicopter configuration are described


in details in the following sections.

a) Pitch (θ)
The blade pitch angle (θ) is the angle between the plane perpendicular to the rotor shaft
and the chord line of a reference station on the blade (Fig. 1).

For a hovering helicopter, angle of incidence (i) is different from θ. As the rotor blade
rotates, a downward velocity (vi) is induced. The resultant velocity VR is a combination
of this induced velocity (vi) and the linear velocity (Ωr) in the plane of rotation at a
distance r from the hub, as shown in Fig. 1. The angle between induced velocity (vi) and
the linear velocity (Ωr) is defined as inflow angle φ and the angle of incidence (i) is
reduced from θ by the inflow angle φ.

It is common knowledge that the lift over an aerofoil is proportional to lift curve slope „a‟
and the angle of incidence i.e., a function of the aerofoil shape and angle of attack,

CL = a.i, CL = lift coeffi


where a is given by linearised theory as
a = 2π if i is in radian
a = 0.11 if i is in degrees

b) Azimuth ()
The helicopter rotor blade moves through 3600 azimuth. The azimuth position  (Fig. 2)
is measured positively in the direction of rotation from its downstream position.

7
c) Change of Pitch
i. Collective change of pitch: Within limits of stall, CL increases with increase in. If
the pitch of all the blades are increased (decreased) simultaneously, the overall lift
and hence thrust increases (decreases). 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. The pilot uses collective pitch lever for this change.

ii. Cyclic change of pitch: With cyclic pitch lever, the pilot can increase the blade pitch
at one azimuth position (A) and decrease it at a diagonally opposite position (B), as
shown in Fig. 3. 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. This cyclic pitch change by using cyclic pitch lever
helps in many ways. It is one of the most effective way of changing the direction of
rotor thrust since this cyclic variation of pitch effectively amounts to tilting of rotor
cone.

Both collective and cyclic pitch change are accomplished by pilot by a swash-plate
system, described later.

d) 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

8
i) 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  = 900) is at a velocity V + r and the retreating blade (at  =
2700) is at V-r. 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. The alternative way is to
introduce flapping blades by use of flapping hinge (Fig. 4), as proposed by Cierva.

With the blades free to flap the moment problem is solved. The blades will move in such
a manner as to seek equilibrium; that is, in such a way as to make the summation of the
moments about the flapping hinge zero.

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 (Fig. 5).
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.

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.

ii) Drag (Lag) Hinge:


The next important hinge is the drag hinge (Fig. 6). 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.

9
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 (Fig. 6). This flexibility makes the net
moment about drag hinge zero.

Both the flapping and drag hinge (in a so-called fully articulated rotor) is shown in Fig.
7c.

iii) 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.

e) Types of Rotors
Three fundamental types of rotors have been developed so far (Fig. 7) :
a. 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.
b. 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.
c. 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.

10
f) Mechanics of Rotor Control
In the case of the conventional single rotor helicopter the control is achieved mainly by
the tilt of the main rotor thrust. Now the problem is to how to tilt the thrust of the main
rotor. One way to do this is to tilt the hub of either a rigid or flapping rotor with respect to
the fuselage. In the normal engine driven helicopter it is mechanically awkward to tilt the
hub since the hub is a rotating structure to which large torque loads are applied.

The most common way of achieving this is by means of a swash-plate system (Fig. 8).
This system consists of two parts: one rotating and the other fixed. This system provides
change in both collective and cyclic pitch.

i.) Collective pitch change: A collective pitch change is applied by raising the fixed
swash plate vertically, which raises the moving swash plate through the same distance
thus ensuring that pitch of all blades changes by the same amount. This change in pitch is
independent of azimuthal position.

ii.) Cyclic pitch change: If the fixed plate is tilted angularly, the moving swash plate is
also titled by the same value. This increases the pitch of one blade in one azimuthal
position and the pitch of the blade at diametrically opposite position will decrease by
same amount.

So this swash plate arrangement when displaced vertically up and down provides
collective pitch change and when titled angularly it provides for cyclic pitch change.

g) Limits of Helicopter Operation : Stall and Compressibility:


Of all, the aerodynamic characteristic peculiar to the helicopter, retreating blade stall is
perhaps the most interesting. Whereas the stall of the wing limits low speed
characteristics of the airplane, stall of the retreating rotor blade imposes a limitation on
high speed capabilities of the helicopter.

11
During forward flight the rotor blade encounters a velocity differential between the
advancing and retreating blades necessitating a change in angle of attack with azimuth
and correction for the resulting dissymmetry of lift. At some particular tip speed ratio, the
retreating blade will reach an angle of attack at which aerofoil will stall. The stall may
begin on any portion of the retreating region depending on the aerodynamic
characteristics of the rotor system. Experience has indicated, however, that for most
conventional rotors, the stall generally begins at the tip at an azimuth of 2700 to 3000 and
propagates deeper into disc with increasing tip speed ratio ( = V cos  / ΩR), as shown
by the shaded are in Fig. 9.

Another limit somewhat similar to stall but on the advancing side may be set by
compressibility effects. The tip of the blade on the advancing side will at high forward
speeds reach the critical Mach number of the blade, causing just like stall, a loss of lift
and an increase in blade drag. Usually stall and compressibility will cause severe
vibrations and excessive oscillatory blade loads. Stall also results in sluggish control and
sometimes a nose up and a slight rolling tendency.

When stall is encountered it may be quickly eliminated by one or a combination of the


following:
1. Decrease airspeed
2. Reduce main rotor pitch
3. Decrease severity of maneuver
4. Increase rpm, unless compressibility effects makes it impractical

Blade stalling can not be avoided entirely, but there are several ways in which it can be
delayed. Some of these methods are:
1. Twisting the blades towards tip
2. Reducing parasite drag, thereby reducing mean lift coefficient and tip angle attack
3. Increase the solidity

12
h) Rotor Blades:
The rotor blades are about 15 to 20 times as long as they are wide. Blades vary in
both planform (Fig. 10) and twist. Best blades from an aerodynamic standpoint
incorporate both twist and taper. Blades re of the following types:

1. All wood blades: are used frequently. They are usually built up from laminations of
several woods, heavier woods being used in the forward section and lightwood such
as Balsa being used in the rear ward portion. Such blades are relatively simple to
fabricate, especially if built with rectangular plan form and constant thickness.
Surfaces are aerodynamically clean and true to contour. However, such blades are
heavy, subject to moisture and deterioration.
2. Metal blades: are being developed at the present time by most manufacturers.
Blades can be built from pieces of sheet metal. It is probably safe to say that all
metal blades will eventually become standard for helicopter rotors.
3. Fabric covered blades: Most early rotor blades employed this type of construction.
The primary structural member of such fabric-covered blades consists of a steel
spar, which is usually step-tapered. Spars are drawn as one continuous tube. The
ribs are usually cut from plywood and are fastened to the spar by metal collars. The
leading edge is built up with solid wood – often with a metal strip to keep the blade
center of gravity forward. The entire blade is covered with fabric. The disadvantage
is that it is difficult to avoid surface irregularities and fabric distortions in flight.
4. Plywood covered blade: Most of the objectionable features of the fabric covered
blades can be overcome by using the same basic structure and covering the entire
blade with thin plywood. However, such blades require careful hardwork, do not
lend themselves to quantity production and are nor weatherproof.

i) Rotor Aerofoil Section:


NANA 0012 aerofoil was selected for early helicopter rotor application and was used
almost exclusively from 1937 to 1977. Alternate aerofoils were hardly considered during
this period because aerodynamic problems were secondary to many structural and
mechanical problems related to flight controls, power systems and structural life.

13
Later, efforts have been made to define more effective aerofoil sections for helicopter
rotors. This has resulted in designing aerofoils tailored to optimize hover, maneuver and
high speed performance simultaneously.

The aerofoil in the helicopter blade is subjected to the most varied and adverse type of
free conditions as it swings around the azimuth. The profiles of an aircraft wing are never
placed in such hostile atmosphere and relatively speaking have an easy life. The
optimization of a helicopter rotor sections is a more difficult task.

First of all, the rotor in hover has uniform flow conditions at all azimuth positions. It
mainly operates up to Mach numbers of 0.5 to 0.6 at the tip gradually reducing towards
the root. The hovering demands a high lift/drag ratio and low pitching moment
coefficient.

In forward flight, the conditions are entirely different. At low altitudes in the higher
ranges of velocity the Mach numbers are quite high at the tip of the advancing blade with
blade sections at low values of CL of 0 to 0.3. On the other hand, the aerofoil on the
retreating blade operates at high values of CL nearing or at stall at Mach numbers of 0.3
to 0.35. At the same time, at the inboard there is a region of reversed flow as well.

In summing up, the aerofoil of the helicopter rotors must show both favorable low speed
and high speed characteristics.

Favorable low speed characteristics include:


high CL at M = 0.5, low CD , low CMa/c ,high CL/CD

Favorable high speed characteristics include:


shock free flow, high MD, low CD, low CMa/c

14
Design requirement for new advanced aerofoils in tabular form :
Flight condition Operating condition Specification
Hover M = 0.6 and CL = 0.65 CL/CD = 72
CM  0.02
1 4c

High Speed Advancing blade: M > 0.85


CL = 0 to 0.03 Shock free flow
Retreating blade: CD < 0.013
CLmax > 1.2 M = 0.3 to 0.35
No separation
Maneuver M = 0.5 CLmax > 1.35
Shock free flow
General 2D Test Condition Re = 5  106

The aerofoil designed must conform to the two transonic design requirements (high speed
flight and maneuver) which simultaneously satisfying stringent subsonic (hover)
requirements as mentioned in the above table.

There are two possibilities in the design process. One is to calculate a suitable shock free
shape for a high CL at Mach number 0.5 and then modify parts of such a basic aerofoil to
optimize towards hover. The other possibility is to start from low speed and optimize
towards high speed and maneuver.

The procedure usually adopted is to start from the high speed side. This is because the
high speed condition will determine a much larger part of the aerofoil counter. Also,
shaping for transonic shock free flow is a very delicate matter and one would certainly
like to leave this designing by some method as much as possible.

Two factors have helped in the designing of “advanced” aerofoils for helicopter rotor.
Firstly, the concept of super-critical aerofoil helped considerably in obtaining suitable
high speed characteristics. The initial stimulus for developing aerofoils with favorable

15
transonic characteristics was given by Piercy. That shock free transonic flow is a real
possibility was proved experimentally by Piercy. He found that shockwaves can be
reduced in strength and even eliminated by designing for a “peaky” type of pressure
distribution.

The first “advanced” aerofoil for helicopter rotor is due to Wortmann who applied the
“peaky” principle to improve the transonic characteristics. Since the Kemp & Piercy et al
have developed other aerofoils designed exclusively for helicopter rotor.

Secondly, though the flow through the rotor is of three-dimensional and unsteady nature,
it has been verified that the performance of a rotor depends strongly on the two-
dimensional steady characteristics of the rotor profiles. In other words, the performance
of a helicopter rotor in a given flight condition can be improved by improving the
characteristics of the rotor aerofoil selected in the two-dimensional steady flow condition.

Overall Design Features of Conventional Helicopter Rotor

1. Profile used : NACA0012, NACA23012, New advanced aerofoils


2. Thickness ratio : 9% - 18%
3. Disc loading : 8 to 48 kg/m2 depending on the type of helicopters
Majority are loaded between 10-20 kg/m2
4. No. of blades : 2 to 6
5. Plan form : Usually rectangular, sometimes trapezoidal blades are
used with taper ratio between 0.5 and 0.7
6. Twist : From root to tip usually between -50 to –120,
-80 to –100 mostly used.
7. Collective pitch at 0.75 R : 60 to 120 in powered flight, 00 to 30 in autorotation
8. Tip speeds : Between 150 to 220 m/sec
9. Mach number at blade tip : 0.92 – 0.97 achieved presently
10. Main operating CL : 0.4 to 0.6
11. Power loading : 2 to 7 kg/hp

16
Tail Rotor
The tail rotor is primarily to counteract the anti-torque due to engine torque. In the
absence of the tail rotor, the helicopter would tend to spin against the main rotor. Thus, it
is always necessary to have an anti-torque device in the form of a tail rotor, situated at a
distance from the center of gravity providing a convenient moment arm for a single rotor
helicopter.

In the conventional tail rotor the working is similar to the main rotor except that it is
much smaller in size. The tail rotor has no drag hinges but only flap hinges.

In the case of twin rotors contra-rotating coaxial main rotors cancel out the torque of each
other. Such helicopter does not need tail rotors.

17
Ω
ω

R
dr

i VR
vi
θ φ
Ωr

Fig. 1 Blade pitch (θ)

18
Front of helicopter

 =1800


 =270 0
 =900

 = 00

Fig. 2 Azimuth position (ψ)

Helicopter motion
θ

B A

Retreating Advancing
blade blade
θ

Fig. 3 Advancing (ψ =900) and retreating (ψ =2700) blades

19

(a) Flapping hinge (without offset)

a

(b) Flapping hinge (with offset a)

Fig. 4 Flapping hinge (with and without offset)

20
VR 

r
V+r
(a) Advancing blade (flapping up reduces angle of incidence)

V-r 

r VR

(b) Retreating blade (flapping down increases angle of incidence)

Fig. 5 Effect of flapping up and down on advancing and retreating blades

21
e

(a) Drag hinge with offset e

(b) Lead-lag motion due to drag hinge

Fig. 6 Drag hinge and its effect

22
(a) Two-bladed rigid rotor (no hinge)

(b) Teetering or see-saw rotor (flapping hinge only)

(c) Fully articulated rotor (flapping and drag hinge)

Fig. 7 Various types of rotors

23
Fig. 8 Swash plate system used in conventional single-rotor helicopter

24
=1800
V=70 mph
=0.23

=900
=270 0

=00

V=70 mph
=0.27
Reversed
flow

=900

V=70 mph
=0.27
Reversed
flow

=900

Angle of attack contour plot Propagation of stall

Fig. 9 Propagation of stall with increase in forward speed of helicopter

25
Fig. 10 Various rotor blade planform shapes

26
Chapter – 2
Performance Analysis of Helicopter in Hover & Vertical Climb
Using Momentum Theory

2.1 Introduction
Various motions of a helicopter can be, broadly, classified as :
a) Hover
b) Vertical Climb
c) Forward motion
d) Vertical Descend
e) Maneuver
The early development of the theory for helicopter motion in hover followed two
independent lines of thought:
i) Momentum theory
ii) Blade element 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.

Notations
(i) Rotor
R = radius of rotor (m)
c = chord of the blade (m)
r = span wise distance of a section from center of rotation (m)
r = r/R
b = number of blades
a = flapping hinge offset
 = pitch of the blade section
0.75R = collective pitch at 0.75 R

27
T = twist (linear) of the blade in degrees
 = inflow angle at the blade element
Ω = angular speed of the blade (rad/sec)
i = incidence of the blade section
M = Mach number at the blade section
CL = coefficient of lift of the blade profile [f (i, M)]
CD = coefficient of drag of the blade profile [f (i, M)]
U = tip speed of rotor (m/sec)
Vi = vertically downward air velocity induced at the rotor disc (m/sec)
Vi∞ = vertically downward air velocity induced at infinity downstream (m/sec)
VR = resultant air velocity at the blade profile (m/sec)
dL = elemental lift
dD = elemental drag
dT = elemental thrust
dFx = elemental inplane force
dQ = elemental torque
w = swirl or rotational speed at the blade element (m/sec)
 = solidity ratio (bc/R)
T = thrust of the rotor (N)
P = power absorbed by the main rotor
 = traction coefficient (=2T/4U2R2)
S = rotor disc area (m2)
S = area at infinity downstream

k = slipstream contraction ratio (= S / S )

(ii) Flight condition


Z = attitude (m)
 = density of air (kg/m3)
m = mass of helicopter (kg)
VZ = rate of climb (m/sec)

28
2.2 Momentum Theory for Vertical Climb
The momentum theory, started by Rankine and further developed by Froude and Betz,
stems from Newton‟s second law of motion F = ma. Although this theory does not
consider the geometry of blades, it results in the general, higher than the speed with
which airscrew advances in air. The increase in velocity of the air from its initial velocity
VZ to its value at the airscrew disk is called the induced or downwash velocity and is
denoted by Vi (Fig. 1). The thrust developed is then equal to the mass of air passing
through the disk in unit time, multiplied by the total increase in velocity caused by the
action of the airscrew.
If  is the air density, S the disc area, the mass flow rate per unit time through the disc
equals  s (VZ  Vi ) . Equating thrust to the change in momentum gives

T  S  (VZ  Vi ) Vi (1)

Condition of continuity of an incompressible flow gives the relationship


S (VZ  Vi )  S  (VZ  Vi ) (2)

Where S and S are the areas of airscrew and wake respectively. Therefore, Eq. (1) may
also be written as
T  S  (VZ  Vi ) Vi (3)

Now this work done by the thrust of the airscrew on the air per unit time is T .(VZ  Vi ) .

This work must be equal to the increase of Kinetic energy of the slipstream per unit time.
This gives
1
T (VZ  Vi )  S  (VZ  Vi ) (VZ  Vi )2 VZ2 
2
Substituting for T from equation (1)
1
S  (VZ  Vi ) Vi (VZ  Vi )  S  (VZ  Vi ) Vi2  2ViVZ 
2
1 Vi2  2Vi VZ 1
or VZ  Vi   Vi  2VZ  (4)
2 Vi 2
Eq. (4) reveals some interesting feature. The axial velocity (VZ + Vi) may be shown as
1
VZ  Vi  VZ   Vz  Vi   (5)
2

29
This shows that the axial velocity through the disc is the average velocity of upstream
fluid VZ and downstream velocity in the wake (VZ + Vi).
Secondly, Eq. (4) also gives directly
1
Vi  Vi or Vi   2Vi (6)
2
This states that induced velocity at the disc is one-half of total increase in velocity
imparted to the air column.

2.2.1 Momentum Theory for Hover


Putting VZ = 0 for hover in Eq. (1) gives
T  S  (VZ  Vi ) Vi

or, T  S  Vi Vi (7)

Using Vi  2Vi (Fig. 2) and S   R 2

T    R2 Vi .2Vi  2   R2Vi 2 (8)

or Vi  T / 2   R 2 (9)

In hover total thrust supports the weight so that T = W and

W DL
Vi   (10)
2R 2
2

1 2 DL
and Vi  Vi  (11)
2 
where DL is the “disk loading” equal to the helicopter weight divided by the disc area
(analogus to “wing loading” for a fixed wing aircraft). Variation of Vi∞ with DL is plotted
in Fig. 3.

This simple relationship illustrates that rotor thrust in hover may be increased by
a. higher density (low attitude)
b. larger disc area (greater rotor diameter)
c. higher downwash velocities (produced by higher collective pitch setting and/or
higher rpm.

30
2.2.2 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.

2.2.3 Rotor Efficiency :


Propeller Criterion of efficiency is given by

useful power T VZ


  (12)
total power P
In hover, it is not possible to apply this condition. Although power P is expanded in
producing thrust T, the transnational velocity VZ is zero, thus  is always zero.
Obviously, lifting rotor needs some other standard of efficiency whereby its lifting ability
may be judged.

A very reasonable way to estimate the efficiency of a lifting rotor is to compare the actual
power required to provide a given thrust with the minimum possible power required to
produce that thrust (i.e., using an “ideal rotor”).

31
Consequently, the criterion of rotor hovering efficiency may be defined as
ideal power required to hover T Vi
M   (13)
actual power required to hover P
where M is called the rotor figure of merit.

Using Eq. (4), Eq. (7) gives

T T W W
M  
P 2 R 2
P 2   R2

DL  PL  power loading  W / P
 PL  DL  disc loading  W /  R 2 
2 
For an ideal rotor, M is 1. A value of 0.75 is considered to be good for a rotor while M =
0.5 indicates poor performance (Fig. 4).

2.2.4 Further Development of Momentum Theory :


It is necessary to modify some of the simplifying assumptions made in the original
theory. It need not be assumed that air flow is uniform all over the disc. Considering an
elemental annular area, elemental thrust may be written as
d T  2 r d r  VZ  Vi  Vi  (14)

Integration gives
R
T   (VZ  Vi ) Vi   2  r dr (15)
0

Validity of Eq. (15) has not been established.

32
Chapter – 3
Performance Analysis of Helicopter in Hover & Vertical Climb
Using Blade Element Theory

3.1 Introduction
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.

3.1 Blade Element Theory for Vertical Climb


Inflow angle (Fig. 5) is given by
VZ  Vi
tan   (1)
r w
and the resultant velocity is given by
VR2  ( r  w)2  (Vz  Vi )2 (2)

Elemental lift and drag can be written as


1
dL   c VR2 CL dr (3)
2
1
dD   c VR2 CD dr (4)
2
where c is the chord.

33
Resolving in the plane of the disc (Fig. 5)
1
dT   c VR2  CL cos   CD sin   dr (5)
2
1
dFx   c VR2  CL sin   CD cos   dr (6)
2
The total thrust can be calculated by
R
dT
T   dr dr
0

R
1
 cVR2  CL cos   CD sin   dr
2 0

for b number of blades


R
b
T    cVR2  CL cos   CD sin   dr
20

3.2 Blade Element Theory for Hover


Elemental lift, obtained in equation (3) is
1
dL   c VR2 CL dr (1)
2
For purposes of simplification, it can be assumed
sin  =
cos  = 1 (2)
VR =Ωr
The blade element lift coefficient may be expressed as
CL = a i = a ( -) (3)
where a = slope of the lift curve slope
With the aid of equations (2) and (3), eq. (1) becomes
1
dT  dL   (  r ) 2 a (   ) c d r (4)
2
For (b) number of blades
1
dT  b  (  r ) 2 a (   ) c d r (5)
2

34
For simplicity in integration it may be assumed that the pitch angle of a blade element
will vary with its radial position r as
R
  t (6)
r
where t is the pitch angle at tip (Fig. 6). This pitch distribution results in a uniform
inflow distribution along the blade span. Such a distribution is therefore labeled ideal
twist, because it yields the minimum induced loss for a given thrust.

The ideal twist results in the following variation on inflow angle


R
  t (7)
r
where t is the inflow angle at tip.
After substituting eq. (6) and eq. (7) in equation (5)
1 R
dT  b  (  r )2 a ( t   t ) c d r (8)
2 r
Integrating eq. (8) over the blade radius, assuming blade c to be constant, the thrust of the
rotor is
b R3
T  2 a ( t   t ) c (9)
2 2
Thrust coefficient may be defined as

CT  T  R 2  ( R )2 (10)

a bc
This gives CT  ( t   t ) (11)
4R

The term solidity may now be introduced. The solidity, of a rotor having rectangular
blades may be defined as the total blade area to the rotor disc area. Thus
bc R bc
   (12)
R 2
R
Expression for CT becomes

CT  a ( t   t ) (13)
4

35
In order to use Eq. (13) easily, it is necessary to replace t by parameters that are known
or easily determined. This is done as follows:
From momentum theory

Vi  T 2   R 2  W 2   R 2 (14)

By definition  t  Vi / R (15)

W
or, t  (16)
2   2 R 4

Thrust, for a helicopter rotor in hover, can be determined using equation (13) together
with equation (16).

36
Chapter – 4
Performance Analysis of Helicopter in Hover
Using Combined Theory

4.1 Introduction
Performance analysis of a helicopter in hover and vertical climbing may be done very
conveniently using a combination of momentum and blade element theory. However,
hover and vertical climb treated separately. Analysis for vertical climb is given in the
next chapter.

4.2 Combined Theory for Hover


For hover (VZ = 0), momentum theory gives an expression for element thrust as a annulus
of radius r (Fig. 7) as
dT  2 r  dr Vi Vi (1)

Continuity equation for hover can be written as


 S Vi   SVi giving

Vi S 1  R 1 
   giving R  2  (2)
Vi S 2  

This above expression may not be valid for a helicopter in hover, particularly for higher
thrust loading. It can be written as
Vi S
  k2 (3)
Vi S
[Slip stream contraction factor k is to be evaluated using an iterative technique, starting
with an initial value 0.707].
Blade momentum theory gives an expression for thrust as
1
dT   cVR2 (CL cos   CD sin  ) dr (4)
2
where VR2  (  r  w)2  Vi 2 VZ  0 for hover  (5)

37
Vi
and tan   (6)
r w

The term solidity () of a rotor having rectangular blades may be defined as the ratio of
the total blade area to rotor disc area
  bcR /  R2  bc /  R
Equating eq. (1) and (4) for (b) number of blades
1
dT  2  r dr  Vi Vi  b .  CVR2 (CL cos   CD sin  ) dr
2
Using equations (3), (5), (6) this leads to

R  2 tan  sin 2
k  (7)
r 2 CL cos   CD sin 

Proceeding in the same manner and equating the elemental torque from the momentum
and blade element theory
b
dQ  2  r dr  Vi 2 wr   c (CL sin   CD cos  ) r VR2 dr (8)
2
This gives after simplification
R  Vi
 (9)
r 8 (CL sin   CD cos  ) VR2

After further simplification with Vi from eq. (6) and VR from eq. (5)

R  sin 2  w
 (10)
r 4 (CL sin   CD cos  ) ( r  w)

bc
or, 4  wrVi  (CL sin   CD cos  ) VR2
2

w Vi bc R  R
or,  
(CL sin   CD cos  ) VR2
8 r R 8 r

38
Solving for w gives

r
w (11)
 4 r sin 2  
1  
 R  (CL sin   CD cos  ) 
The thrust is given by

R
b
T   cVR2 (CL cos   CD sin  ) dr (12)
0
2

Power is given by

R
b
P  Q    cVR2 (CL sin   CD cos  ) r dr (13)
0
2

4.3 Method of Computation :


Pitch 0 (or ө0.75R ) is given :
1. Divide the blade into number of equal parts
2. Assume a contraction ratio factor k to start with (initial value can be taken as
k  1/ 2  0.707 )
3. Take one station and calculate local pitch using the formula  (.75 R )  (r  .75)T

4. Assume a value of  to start with (initial value can be taken as  = 0)


5. Calculate i, i =  - 
r
6. Local Mach number M  (T in 0C )
20.1 T  273

7. Find CL, CD for i and Mach number M from 2D polar of given profile
8. For the assumed value of k (contraction ratio) determine  from the equation (7)

R  2 tan  sin 2 
k  (7)
r 2 CL cos   CD sin 

39
9. To get a converged value of , iterate steps between steps (5) and (8) where new
values of  are assumed each time.
10. At each station calculate the rotational component of induced velocity from
equation
r
w
 4 r sin 2  
1  
 R  (CL sin   CD cos  ) 

11. At each station calculate the axial induced velocity


Vi  tan  (r  w)

12. Calculate the resultant velocity


VR2  (r  w)2 (1 tan 2  )

1
13. At each station calculate the elemental thrust dT   cVR2 (CL cos   CD sin  ) dr
2
14. Repeat steps (3) to (13)
R
dT
15. Calculate the total thrust by integrating T  b
R0

dr
dr

R
1
b  2  cV (CL cos   CD sin  ) dr
2
R
R0

16. Obtain new value of k from the following figure


17. Compare the new value of k with assumed (old) value of k and repeat steps (2) to
(16) till desired accuracy in achieved in k.
18. Calculate the total power
R
1
P  2  c (C
R0
L sin   CD cos  ) rVR2 dr

40
[Proof: for Eq.(7)

From Eq. (7), replacing the value of r,  and K gives

R  2 R bc Vi
L. H.S  K 
r 2 bcVR (CL cos   CD sin  ) 2  R Vi
2

4  Vi Vi

R 4  Vi Vi bc Vi

bcV (CL cos   CD sin  ) 2  R Vi
R
2

2Vi 2
 2
VR (CL cos   CD sin  )

sin 
tan  sin 2  cos 2sin  cos 
R. H . S  
CL cos   CD sin   CL cos   CD sin 

2sin 2  1  cos 2 
 
CL cos   CD sin  CL cos   CD sin 


1  11 tan
2

tan 2  2 tan 2 
 
CL cos   CD sin  (1  tan 2  ) (CL cos   CD sin  )

 
2
1 Vi
r w 2Vi 2
 
1 
    C cos   CD sin     r  w 2  (CL cos   CD sin  )
Vi 2

 r w L  

2 Vi 2
 2 ]
VR  CL cos   CD sin  

41
Chapter – 5
Performance Analysis of Helicopter in Vertical Climb
Using Combined Theory

5.1 Combined Theory for Vertical Climb


In vertical climb, the contraction ratio is not only a function of thrust loading but of the
rate of climb as well. As the rotor climbs from hover, the induced velocity Vi decreases
as VZ increases. However, for all practical purposes it is assumed that the induced
velocity Vi at downstream infinity is twice that at rotor disc.

Elemental thrust given by momentum theory for an annulus of width dr and at a radial
distance r
dT  2 r  VZ  Vi 2Vi dr (1)

Flow conditions on a blade element will be similar as in the case of hovering except that
the axial inflow velocity will be VZ + Vi instead of Vi.

The inflow angle  is given by

Vz  Vi
tan   (2)
r  w
Resultant velocity VR is given by

VR2  ( r  w)2  (VZ  Vi )2 (3)

or, VR2  ( r  w)2  (1 tan 2  ) (4)

Elemental forces in the plane of the disc are, as for hover,

42
1
dT   cVR2  CL cos   CD sin   dr
2

1
dT   cVR2  CL sin   CD cos   dr
2
Equating the thrust from momentum and blade element theory (for b number of blades)
b
2  r dr  VZ  Vi 2Vi   cVR2  CL cos   CD sin   dr
2

R sin 2  [( r  w) tan   VZ ]


or,  (5)
4r ( r  w)[CL cos   CD sin  ]

Similarly, equating elemental torques from blade element and momentum theories:
b
dQ  2  r dr  VZ  Vi 2 wr   c  CL sin   CD cos   rVR2 dr
2
Rearranging,
R  sin 2  w
 (6)
r 4 (CL sin   CD cos  ) (r  w)

r
where, w (7)
 4r sin 2  
1  
 R  (CL sin   CD cos  ) 

The thrust and power are given by same as in hover

R
b
T  c VR2 (CL cos   CD sin  ) dr (8)
R0
2

R
b
P   c r VR2 (CL sin   CD cos  ) dr (9)
R0
2

43
5.2 Method of Computation
1. Divide the blade into number of equal parts
2.  00.75 R  (r  0.75) T

3. Assume 
4. i   
r
5. Mach no. M  (T in 0 c)
20.1 T  27.3

6. For i, M, find CL, CD


r
7. w 
 4 r sin 2  
1  
 R  (CL sin   CD cos  ) 
8. check if equation (5)

R sin 2   r  w tan   VZ 


 is satisfied
4r R  (CL sin   CD cos  )

9. If it is not satisfied, repeat (2) to (10) with new value of .


10. For next station repeat steps (2) to (10).
11. At each station, calculate Vi  tan  (r  w) VZ

12. At each station, calculate VR2  (r  w)2 (1 tan 2  )


R
b
13. T    c VR2 (CL cos   CD sin  ) dr
R0
2
R
b
14. P    c VR2 (CL sin   CD cos  ) r dr
R0
2

44
Chapter – 6
Performance Analysis of Helicopter in Descending Flight

6.1 Introduction
Four different flow states have been identified with descending flight.
(b) normal thrusting state
(c) vortex ring state
(d) autorotative state
(e) windmill brake state

Figs 8-9 give idealised representation of these states for a purely vertical descent

6.2 Normal thrusting state:


For hovering and fairly low rates of descent the velocity of air induced through the
disc (Vi) exceeds the rate of descent (VZ) itself. In this state all flow through the rotor
is downward but not necessarily of equal magnitude because of the non-uniform
condition of speed and angle of attack from root to tip. Thrust is quite steady.

6.3 Vortex ring state:


For descent rates upto 150% of hover, a condition of large variations in thrust is
experienced with accompany increased vibration and tendencies to produce even
higher rates of descent. This condition, known to the pilots, as “Settling with power”
is more formally called vortex ring state and is somewhat like flying in one‟s own
wake. As can be seen from the figure the high rate of descent has overcome the
normal downward induced flow on inner blade sections. The flow is thus upward,
relative to the rotor disc in these areas and downward at outboard.

This produces a secondary vortex ring in addition to normal tip vortex system. The
result of this set of vortices is unsteady turbulent flow over a large area of the disc,
with an accompanying loss of thrust and excessive thrust fluctuations even through

45
power is still being supplied from the engine. Pilots are warned to avoid situations
that create this condition (i.e., steep descent at high rates of descent).

6.4 Autorotative State :


Beyond the vortex ring state, things settle down again in terms of the intensity of the
turbulent, unsteady flow. There is some rate of descent in vertical flight between 150
to 180% of (Vi) hover where no power is required to maintain the rpm. This state of
autorotation is, of course, extremely important in cases of engine failure when one
wishes to produce thrust, equal to weight, in order to effect controlled equilibrium
flight to the ground at reasonable rates of descent. In essence, potential energy is used
at a rate just sufficient to provide the power requirement in vertical flight.

Helicopters are equipped with overriding clutches so that, in the event of power
failure, the rotor will not be restrained by the engine but will be free to rotate.
Immediately, after a power failure, the pilot must “dump” his collective pitch within 2
to 3 seconds. With decreased collective pitch, rotor will auto rotate as the helicopter
begins to descend; that is the aerodynamic forces on the rotor will cause it to rotate
even though no mechanical torque is present.

6.5 Windmill brake state:


If the rotor descends at rates in excess of 180% of (Vi)hover, it is necessary to brake the
system in order to maintain RPM. In this state, all the flow is “up” relative to the rotor
and energy may be extracted from the system. This is the state in which windmills
operate, extracting energy from the flow of air past them. This is not a normal
operating state of any helicopter.

46
Chapter – 7
Helicopter Control

7.1 Introduction
Control methods are discussed, first from the over-all point of view of the forces and
moments applied to the helicopter and second, from the point of view of the levers which
the pilot operates.

7.2 Control Requirements:


To control completely the position and attitude of a body in space requires control of the
forces and moments about the three axes (Fig. 10).

This involves six independent controls. For example, if the body drifts to a side, a force
may be exerted to return it to its original position. If it rolls, a moment may be exerted to
right it again. However, it will be exceedingly difficult for a human pilot to coordinate six
independent controls. Fortunately, it is possible to reduce this number by coupling
together independent controls, although such couplings involve some sacrifice of
complete freedom of control of position and attitude (in space).

The pilot of a helicopter does demand the ability to produce moments about all axes in
order to right himself, when disturbed by a gust. He does not, however, demand that he
be able to produce moments without producing an accompanying force. For example, if a
pitching moment is produced along with an accompanying force in longitudinal direction.
By this coupling of pitching moment with longitudinal force, the necessity for one of the
independent controls is implemented.

Actually, only four independent controls are adequate for the helicopter:
(a) Vertical control
(b) Directional control
(c) Longitudinal control
(d) Lateral control

47
a) Vertical control: Vertical control is necessary to fix the position of the helicopter
in the vertical direction, i.e., providing means for climbing and descending flight. It is
achieved by increasing or decreasing the collective pitch of the main rotor. By increasing
collective pitch we mean that the pitch of all the blades has been increased by the same
amount and that pitch is independent of azimuthal position of the blade.
b) Directional control: Directional control fixes the attitude of the helicopter in
rotation about the vertical axis, permitting the pilot to point the ship in any horizontal
direction. This is achieved by either by changing the pitch (thrust) of the tail rotor in
conventional single rotor helicopter or by obtaining differential torque in case of twin
main rotor helicopter.
c) Longitudinal control: Longitudinal control involves the application of both
moments and forces. Pitching moments are coupled with longitudinal force. When the
pilot operates the longitudinal control, a pitching moment is produced about the
helicopter C.G. which tilts the helicopter in forward direction. As a consequence of the
tilt, a component of the rotor thrust vector acts in the direction of the tilt. The application
of longitudinal control has therefore resulted in a forward tilt and forward motion of the
helicopter. In longitudinal control, thus, moment is coupled with force.
d) Lateral control: Lateral control is identical in nature to longitudinal control.
Lateral control results in rolling moment as well as sideward motion of the helicopter.

7.3 Pilot’s Control


There are generally, four control levels to be operated by the pilot. They are:
(a) The control stick (cyclic pitch lever)
(b) Pitch lever (collective pitch lever)
(c) Pedals
(d) Throttle

a) The Control Stick: Control stick is located in first of the pilot and is operated by his
right hand. In fact it is comparable to the stick used in fixed wing aircraft and is used for

48
longitudinal and lateral control. It controls the cyclic pitch of the main rotor. It can be
displaced fore and aft and sideways as well as combination of these two motions.

In the helicopter, the pilot pushes the stick in the direction he wishes to go – forward,
sideward or even backward. For example, if the pilot wants to go forward he moves the
stick in forward direction. Similarly, if he wants to go towards his right, he moves his
stick towards his right.

Forward motion: To go forward the rotor cone must be tilted forward. To tilt to rotor
cone forward, the rotor blade must flap up at  = 00 and flap down at  = 1800. Thus,
with the blade flapping high over the tail and low over the nose, the rotor disc is
effectively tilted forward.

Sideward motion: Similarly, if the pilot wants to go to right, blade should flap up at
 = 2700 and flap down at  = 900 effectively tilting the rotor thrust towards right.

Phase-Lag: It is worth noting that when the pitch of a blade is increased the blade
does not flap up instantaneously. There is a phase lag between the application of the force
on the rotor and the ensuing displacement due to the inertia of the blade. The blade can
not deviate immediately from its path of rotation but does so 900 later. Therefore, to
create forward motion, blade pitch would be increased at  = 2700 and decreased at  =
900 and for sideward motion at  = 1800 and  = 00 so that the desired flapping up and
down occurs 900 later.

b) Collective Pitch Lever: Collective pitch lever is situated on the left hand side of
the pilot and is operated by his left hand. It is used for up and down motion of the and for
adjustments as required forward flight. The pitch lever operates the collective pitch of the
main rotor. When pilot wants to go up he moves the pitch lever upwards, which increases
the collective pitch of the main rotor and the helicopter starts climbing because of
increased thrust.

49
c) Pedals: The pedals are situated at the floor in front of the pilot. They are two
numbers, left and right, and they are operated by the pilot‟s feet. They move
differentially, i.e., when left pedal is pushed, the right pedal comes out towards pilot.
Pedals are used for directional control of the helicopter. To point the helicopter towards
right, the pilot pushes the right pedal, to the left, the left pedal.

The pedals are connected to the pitch of the tail rotor. Under equilibrium, the antitorque
of the main rotor is balanced by the thrust of the tail rotor. When the pilot pushes the
right pedal, the pitch of the tail rotor is increased which gives a left force at the tail rotor
which in turn produces a nose right moment.

d) Throttle: The throttle controls the power output from the engine. It is usually
located near or on the pitch lever. Throttle adjustments are made by twisting a grip
located at the top of the pitch lever.

50
Vz

Vz+Vi

Vz+ Vi
S
Fig. 1 Flow through rotor disc in vertical climb

Vz=0

Vi
R
Vi

Vi Vi =2Vi


R
Flow field Velocity variation

Fig. 2 Flow through the rotor disc in hover

51
2 DL
Vi = m/sec

Downwash
Velocity
Vi

1000ft Sea level

DL

Fig. 3 Disc loading characteristics

PL M =1 (ideal rotor)

M = . 75 (good rotor) At sea level

M =. 5
(poor rotor)

DL

Fig. 4 Performance characteristics of a rotor

52
dL

dT

i

dFx  VZ + Vi
r W
dD

Fig. 5 Flow characteristics on a rotor blade element in vertical climb

Vi t Vi  
t
R

Tip section Inboard section

Fig. 6 Variation of pitch and inflow angle at different blade sections in hover

r dr

dT = 2 π dr ρ Vi Vi∞

Fig. 7 Flow through the rotor annulus

53
54
Vi
Vi hover
B
A C
VZ =0

D Vi

1 Hover VZ

Climb Vi

Vi decreases as
Vz increases

-3 -2 -1 0 1 2 3

VZ
Vi hover
Descend Climb

A = Windmill brake state


B = Autorotative states
C = Vortex ring state
D = Normal thrusting state

Fig. 9 Graph of Flow state

55
56

You might also like