You are on page 1of 49

How Airplanes Fly: A Physical Description of Lift

Level 3

Almost everyone today has flown in an airplane. Many ask the simple question "what
makes an airplane fly"? The answer one frequently gets is misleading and often just
plain wrong. We hope that the answers provided here will clarify many
misconceptions about lift and that you will adopt our explanation when explaining lift
to others. We are going to show you that lift is easier to understand if one starts with
Newton rather than Bernoulli. We will also show you that the popular explanation that
most of us were taught is misleading at best and that lift is due to the wing diverting
air down.

Let us start by defining three descriptions of lift commonly used in textbooks and
training manuals. The first we will call the Mathematical Aerodynamics
Description which is used by aeronautical engineers. This description uses complex
mathematics and/or computer simulations to calculate the lift of a wing. These are
design tools which are powerful for computing lift but do not lend themselves to an
intuitive understanding of flight.

The second description we will call the Popular Explanation which is based on the
Bernoulli principle. The primary advantage of this description is that it is easy to
understand and has been taught for many years. Because of its simplicity, it is used to
describe lift in most flight training manuals. The major disadvantage is that it relies on
the "principle of equal transit times" which is wrong. This description focuses on the
shape of the wing and prevents one from understanding such important phenomena as
inverted flight, power, ground effect, and the dependence of lift on the angle of attack
of the wing.

The third description, which we are advocating here, we will call the Physical
Description of lift. This description is based primarily on Newton’s laws. The physical
description is useful for understanding flight, and is accessible to all that are curious.
Little math is needed to yield an estimate of many phenomena associated with flight.
This description gives a clear, intuitive understanding of such phenomena as the
power curve, ground effect, and high-speed stalls. However, unlike the mathematical
aerodynamics description, the physical description has no design or simulation
capabilities.
The popular explanation of lift
Students of physics and aerodynamics are taught that airplanes fly as a result of
Bernoulli’s principle, which says that if air speeds up the pressure is lowered. Thus a
wing generates lift because the air goes faster over the top creating a region of low
pressure, and thus lift. This explanation usually satisfies the curious and few challenge
the conclusions. Some may wonder why the air goes faster over the top of the wing
and this is where the popular explanation of lift falls apart.

In order to explain why the air goes faster over the top of the wing, many have
resorted to the geometric argument that the distance the air must travel is directly
related to its speed. The usual claim is that when the air separates at the leading edge,
the part that goes over the top must converge at the trailing edge with the part that
goes under the bottom. This is the so-called "principle of equal transit times".

As discussed by Gale Craig (Stop Abusing Bernoulli! How Airplanes Really Fly.,
Regenerative Press, Anderson, Indiana, 1997), let us assume that this argument were
true. The average speeds of the air over and under the wing are easily determined
because we can measure the distances and thus the speeds can be calculated. From
Bernoulli’s principle, we can then determine the pressure forces and thus lift. If we do
a simple calculation we would find that in order to generate the required lift for a
typical small airplane, the distance over the top of the wing must be about 50% longer
than under the bottom. Figure 1 shows what such an airfoil would look like. Now,
imagine what a Boeing 747 wing would have to look like!

Fig 1 Shape of wing predicted by principle of equal transit time.

If we look at the wing of a typical small plane, which has a top surface that is 1.5 -
2.5% longer than the bottom, we discover that a Cessna 172 would have to fly at over
400 mph to generate enough lift. Clearly, something in this description of lift is
flawed.

But, who says the separated air must meet at the trailing edge at the same time? Figure
2 shows the airflow over a wing in a simulated wind tunnel. In the simulation, colored
smoke is introduced periodically. One can see that the air that goes over the top of the
wing gets to the trailing edge considerably before the air that goes under the wing. In
fact, close inspection shows that the air going under the wing is slowed down from the
"free-stream" velocity of the air. So much for the principle of equal transit times.

Fig 2 Simulation of the airflow over a wing in a wind tunnel, with colored "smoke" to show
the acceleration and deceleration of the air.

The popular explanation also implies that inverted flight is impossible. It certainly
does not address acrobatic airplanes, with symmetric wings (the top and bottom
surfaces are the same shape), or how a wing adjusts for the great changes in load such
as when pulling out of a dive or in a steep turn?

So, why has the popular explanation prevailed for so long? One answer is that
the Bernoulli principle is easy to understand. There is nothing wrong with the
Bernoulli principle, or with the statement that the air goes faster over the top of the
wing. But, as the above discussion suggests, our understanding is not complete with
this explanation. The problem is that we are missing a vital piece when we apply
Bernoulli’s principle. We can calculate the pressures around the wing if we know the
speed of the air over and under the wing, but how do we determine the speed?

Another fundamental shortcoming of the popular explanation is that it ignores the


work that is done. Lift requires power (which is work per time). As will be seen later,
an understanding of power is key to the understanding of many of the interesting
phenomena of lift.

Newton’s laws and lift


So, how does a wing generate lift? To begin to understand lift we must return to high
school physics and review Newton’s first and third laws. (We will introduce Newton’s
second law a little later.) Newton’s first law states a body at rest will remain at rest,
or a body in motion will continue in straight-line motion unless subjected to an
external applied force. That means, if one sees a bend in the flow of air, or if air
originally at rest is accelerated into motion, there is a force acting on it. Newton’s
third law states that for every action there is an equal and opposite reaction. As an
example, an object sitting on a table exerts a force on the table (its weight) and the
table puts an equal and opposite force on the object to hold it up. In order to generate
lift a wing must do something to the air. What the wing does to the air is the action
while lift is the reaction.

Let’s compare two figures used to show streams of air (streamlines) over a wing. In
figure 3 the air comes straight at the wing, bends around it, and then leaves straight
behind the wing. We have all seen similar pictures, even in flight manuals. But, the air
leaves the wing exactly as it appeared ahead of the wing. There is no net action on the
air so there can be no lift! Figure 4 shows the streamlines, as they should be drawn.
The air passes over the wing and is bent down. The bending of the air is the action.
The reaction is the lift on the wing.

Fig 3 Common depiction of airflow over a wing. This wing has no lift.

Fig 4 True airflow over a wing with lift, showing upwash and downwash.

The wing as a pump


As Newton’s laws suggests, the wing must change something of the air to get lift.
Changes in the air’s momentum will result in forces on the wing. To generate lift a
wing must divert air down; lots of air.
The lift of a wing is equal to the rate of change in momentum of the air it is diverting
down. Momentum is the product of mass and velocity. The lift of a wing is
proportional to the amount of air diverted down per second times the downward
velocity of that air. Its that simple. (Here we have used an alternate form of Newton’s
second law that relates the acceleration of an object to its mass and to the force on it;
F=ma) For more lift the wing can either divert more air (mass) or increase its
downward velocity. This downward velocity behind the wing is called "downwash".
Figure 5 shows how the downwash appears to the pilot (or in a wind tunnel). The
figure also shows how the downwash appears to an observer on the ground watching
the wing go by. To the pilot the air is coming off the wing at roughly the angle of
attack. To the observer on the ground, if he or she could see the air, it would be
coming off the wing almost vertically. The greater the angle of attack, the greater the
vertical velocity. Likewise, for the same angle of attack, the greater the speed of the
wing the greater the vertical velocity. Both the increase in the speed and the increase
of the angle of attack increase the length of the vertical arrow. It is this vertical
velocity that gives the wing lift.

Fig 5 How downwash appears to a pilot and to an observer on the ground.

As stated, an observer on the ground would see the air going almost straight down
behind the plane. This can be demonstrated by observing the tight column of air
behind a propeller, a household fan, or under the rotors of a helicopter; all of which
are rotating wings. If the air were coming off the blades at an angle the air would
produce a cone rather than a tight column. If a plane were to fly over a very large
scale, the scale would register the weight of the plane.

If we estimate that the average vertical component of the downwash of a Cessna 172
traveling at 110 knots to be about 9 knots, then to generate the needed 2,300 lbs of lift
the wing pumps a whopping 2.5 ton/sec of air! In fact, as will be discussed later, this
estimate may be as much as a factor of two too low. The amount of air pumped down
for a Boeing 747 to create lift for its roughly 800,000 pounds takeoff weight is
incredible indeed.

Pumping, or diverting, so much air down is a strong argument against lift being just a
surface effect as implied by the popular explanation. In fact, in order to pump 2.5
ton/sec the wing of the Cessna 172 must accelerate all of the air within 9 feet above
the wing. (Air weighs about 2 pounds per cubic yard at sea level.) Figure 6 illustrates
the effect of the air being diverted down from a wing. A huge hole is punched through
the fog by the downwash from the airplane that has just flown over it.

Fig 6 Downwash and wing vortices in the fog.


(Photographer Paul Bowen, courtesy of Cessna Aircraft, Co.)

So how does a thin wing divert so much air? When the air is bent around the top of
the wing, it pulls on the air above it accelerating that air down, otherwise there would
be voids in the air left above the wing. Air is pulled from above to prevent voids. This
pulling causes the pressure to become lower above the wing. It is the acceleration of
the air above the wing in the downward direction that gives lift. (Why the wing bends
the air with enough force to generate lift will be discussed in the next section.)

As seen in figure 4, a complication in the picture of a wing is the effect of "upwash" at


the leading edge of the wing. As the wing moves along, air is not only diverted down
at the rear of the wing, but air is pulled up at the leading edge. This upwash actually
contributes to negative lift and more air must be diverted down to compensate for it.
This will be discussed later when we consider ground effect.

Normally, one looks at the air flowing over the wing in the frame of reference of the
wing. In other words, to the pilot the air is moving and the wing is standing still. We
have already stated that an observer on the ground would see the air coming off the
wing almost vertically. But what is the air doing above and below the wing? Figure 7
shows an instantaneous snapshot of how air molecules are moving as a wing passes
by. Remember in this figure the air is initially at rest and it is the wing moving. Ahead
of the leading edge, air is moving up (upwash). At the trailing edge, air is diverted
down (downwash). Over the top the air is accelerated towards the trailing edge.
Underneath, the air is accelerated forward slightly, if at all.

Fig 7 Direction of air movement around a wing as seen by an observer on the ground.

In the mathematical aerodynamics description of lift this rotation of the air around the
wing gives rise to the "bound vortex" or "circulation" model. The advent of this
model, and the complicated mathematical manipulations associated with it, leads to
the direct understanding of forces on a wing. But, the mathematics required typically
takes students in aerodynamics some time to master.

One observation that can be made from figure 7 is that the top surface of the wing
does much more to move the air than the bottom. So the top is the more critical
surface. Thus, airplanes can carry external stores, such as drop tanks, under the wings
but not on top where they would interfere with lift. That is also why wing struts under
the wing are common but struts on the top of the wing have been historically rare. A
strut, or any obstruction, on the top of the wing would interfere with the lift.

Air has viscosity


The natural question is "how does the wing divert the air down?" When a moving
fluid, such as air or water, comes into contact with a curved surface it will try to
follow that surface. To demonstrate this effect, hold a water glass horizontally under a
faucet such that a small stream of water just touches the side of the glass. Instead of
flowing straight down, the presence of the glass causes the water to wrap around the
glass as is shown in figure 8. This tendency of fluids to follow a curved surface is
known as the Coanda effect. From Newton’s first law we know that for the fluid to
bend there must be a force acting on it. From Newton’s third law we know that the
fluid must put an equal and opposite force on the object which caused the fluid to
bend.

Fig 8 Coanda effect.

Why should a fluid follow a curved surface? The answer is viscosity; the resistance to
flow which also gives the air a kind of "stickiness". Viscosity in air is very small but it
is enough for the air molecules to want to stick to the surface. At the surface the
relative velocity between the surface and the nearest air molecules is exactly zero.
(That is why one cannot hose the dust off of a car and why there is dust on the
backside of the fans in a wind tunnel.) Just above the surface the fluid has some small
velocity. The farther one goes from the surface the faster the fluid is moving until the
external velocity is reached (note that this occurs in less than an inch). Because the
fluid near the surface has a change in velocity, the fluid flow is bent towards the
surface. Unless the bend is too tight, the fluid will follow the surface. This volume of
air around the wing that appears to be partially stuck to the wing is called the
"boundary layer".

Lift as a function of angle of attack


There are many types of wing: conventional, symmetric, conventional in inverted
flight, the early biplane wings that looked like warped boards, and even the proverbial
"barn door". In all cases, the wing is forcing the air down, or more accurately pulling
air down from above. What each of these wings have in common is an angle of attack
with respect to the oncoming air. It is this angle of attack that is the primary parameter
in determining lift. The inverted wing can be explained by its angle of attack, despite
the apparent contradiction with the popular explanation involving the Bernoulli
principle. A pilot adjusts the angle of attack to adjust the lift for the speed and load.
The popular explanation of lift which focuses on the shape of the wing gives the pilot
only the speed to adjust.

To better understand the role of the angle of attack it is useful to introduce an


"effective" angle of attack, defined such that the angle of the wing to the oncoming air
that gives zero lift is defined to be zero degrees. If one then changes the angle of
attack both up and down one finds that the lift is proportional to the angle. Figure 9
shows the coefficient of lift (lift normalized for the size of the wing) for a typical wing
as a function of the effective angle of attack. A similar lift versus angle of attack
relationship is found for all wings, independent of their design. This is true for the
wing of a 747 or a barn door. The role of the angle of attack is more important than
the details of the airfoil’s shape in understanding lift.

Fig 9 Coefficient of lift versus the effective angle of attack.

Typically, the lift begins to decrease at an angle of attack of about 15 degrees. The
forces necessary to bend the air to such a steep angle are greater than the viscosity of
the air will support, and the air begins to separate from the wing. This separation of
the airflow from the top of the wing is a stall.

The wing as air "scoop"


We now would like to introduce a new mental image of a wing. One is used to
thinking of a wing as a thin blade that slices though the air and develops lift somewhat
by magic. The new image that we would like you to adopt is that of the wing as a
scoop diverting a certain amount of air from the horizontal to roughly the angle of
attack, as depicted in figure 10. The scoop can be pictured as an invisible structure put
on the wing at the factory. The length of the scoop is equal to the length of the wing
and the height is somewhat related to the chord length (distance from the leading edge
of the wing to the trailing edge). The amount of air intercepted by this scoop is
proportional to the speed of the plane and the density of the air, and nothing else.

Fig 10 The wing as a scoop.

As stated before, the lift of a wing is proportional to the amount of air diverted down
times the vertical velocity of that air. As a plane increases speed, the scoop diverted
more air. Since the load on the wing, which is the weight of the plane, does not
increase the vertical speed of the diverted air must be decreased proportionately. Thus,
the angle of attack is reduced to maintain a constant lift. When the plane goes higher,
the air becomes less dense so the scoop diverts less air for the same speed. Thus, to
compensate the angle of attack must be increased. The concepts of this section will be
used to understand lift in a way not possible with the popular explanation.

Lift requires power


When a plane passes overhead the formerly still air ends up with a downward
velocity. Thus, the air is left in motion after the plane leaves. The air has been given
energy. Power is energy, or work, per time. So, lift must require power. This power is
supplied by the airplane’s engine (or by gravity and thermals for a sailplane).

How much power will we need to fly? The power needed for lift is the work (energy)
per unit time and so is proportional to the amount of air diverted down times the
velocity squared of that diverted air. We have already stated that the lift of a wing is
proportional to the amount of air diverted down times the downward velocity of that
air. Thus, the power needed to lift the airplane is proportional to the load (or weight)
times the vertical velocity of the air. If the speed of the plane is doubled the amount of
air diverted down doubles. Thus the angle of attack must be reduced to give a vertical
velocity that is half the original to give the same lift. The power required for lift has
been cut in half. This shows that the power required for lift becomes less as the
airplane's speed increases. In fact, we have shown that this power to create lift is
proportional to one over the speed of the plane.

But, we all know that to go faster (in cruise) we must apply more power. So there
must be more to power than the power required for lift. The power associated with
lift, described above, is often called the "induced" power. Power is also needed to
overcome what is called "parasitic" drag, which is the drag associated with moving
the wheels, struts, antenna, etc. through the air. The energy the airplane imparts to an
air molecule on impact is proportional to the speed squared. The number of molecules
struck per time is proportional to the speed. Thus the parasitic power required to
overcome parasitic drag increases as the speed cubed.

Figure 11 shows the power curves for induced power, parasitic power, and total power
which is the sum of induced power and parasitic power. Again, the induced power
goes as one over the speed and the parasitic power goes as the speed cubed. At low
speed the power requirements of flight are dominated by the induced power. The
slower one flies the less air is diverted and thus the angle of attack must be increased
to maintain lift. Pilots practice flying on the "backside of the power curve" so that
they recognizes that the angle of attack and the power required to stay in the air at
very low speeds are considerable.

Fig 11 Power requirements versus speed.


At cruise, the power requirement is dominated by parasitic power. Since this goes as
the speed cubed an increase in engine size gives one a faster rate of climb but does
little to improve the cruise speed of the plane.

Since we now know how the power requirements vary with speed, we can understand
drag, which is a force. Drag is simply power divided by speed. Figure 12 shows the
induced, parasitic, and total drag as a function of speed. Here the induced drag varies
as one over speed squared and parasitic drag varies as the speed squared. Taking a
look at these curves one can deduce a few things about how airplanes are designed.
Slower airplanes, such as gliders, are designed to minimize induced drag (or induced
power), which dominates at lower speeds. Faster airplanes are more concerned with
parasite drag (or parasitic power).

Fig 12 Drag versus speed.

Wing efficiency
At cruise, a non-negligible amount of the drag of a modern wing is induced drag.
Parasitic drag, which dominates at cruise, of a Boeing 747 wing is only equivalent to
that of a 1/2-inch cable of the same length. One might ask what effects the efficiency
of a wing. We saw that the induced power of a wing is proportional to the vertical
velocity of the air. If the length of a wing were to be doubled, the size of our scoop
would also double, diverting twice as much air. So, for the same lift the vertical
velocity (and thus the angle of attack) would have to be halved. Since the induced
power is proportional to the vertical velocity of the air, it too is reduced by half. Thus,
the lifting efficiency of a wing is proportional to one over the length of the wing. The
longer the wing the less induced power required to produce the same lift, though this
is achieved with an increase in parasitic drag. Low speed airplanes are affected more
by induced drag than fast airplanes and so have longer wings. That is why sailplanes,
which fly at low speeds, have such long wings. High-speed fighters, on the other
hand, feel the effects of parasite drag more than our low speed trainers. Therefore, fast
airplanes have shorter wings to lower parasite drag.

There is a misconception by some that lift does not require power. This comes from
aeronautics in the study of the idealized theory of wing sections (airfoils). When
dealing with an airfoil, the picture is actually that of a wing with infinite span. Since
we have seen that the power necessary for lift is proportional to one over the length of
the wing, a wing of infinite span does not require power for lift. If lift did not require
power airplanes would have the same range full as they do empty, and helicopters
could hover at any altitude and load. Best of all, propellers (which are rotating wings)
would not require power to produce thrust. Unfortunately, we live in the real world
where both lift and propulsion require power.

Power and wing loading


Let us now consider the relationship between wing loading and power. Does it take
more power to fly more passengers and cargo? And, does loading affect stall speed?
At a constant speed, if the wing loading is increased the vertical velocity must be
increased to compensate. This is done by increasing the angle of attack. If the total
weight of the airplane were doubled (say, in a 2g turn) the vertical velocity of the air
is doubled to compensate for the increased wing loading. The induced power is
proportional to the load times the vertical velocity of the diverted air, which have both
doubled. Thus the induced power requirement has increased by a factor of four! The
same thing would be true if the airplane’s weight were doubled by adding more fuel,
etc.

One way to measure the total power is to look at the rate of fuel consumption. Figure
13 shows the fuel consumption versus gross weight for a large transport airplane
traveling at a constant speed (obtained from actual data). Since the speed is constant
the change in fuel consumption is due to the change in induced power. The data are
fitted by a constant (parasitic power) and a term that goes as the load squared. This
second term is just what was predicted in our Newtonian discussion of the effect of
load on induced power.
Fig 13 Fuel consumption versus load for a large transport airplane
traveling at a constant speed.

The increase in the angle of attack with increased load has a downside other than just
the need for more power. As shown in figure 9 a wing will eventually stall when the
air can no longer follow the upper surface. That is, when the critical angle is reached.
Figure 14 shows the angle of attack as a function of airspeed for a fixed load and for a
2-g turn. The angle of attack at which the plane stalls is constant and is not a function
of wing loading. The stall speed increases as the square root of the load. Thus,
increasing the load in a 2-g turn increases the speed at which the wing will stall by
40%. An increase in altitude will further increase the angle of attack in a 2-g turn.
This is why pilots practice "accelerated stalls" which illustrates that an airplane can
stall at any speed. For any speed there is a load that will induce a stall.
Fig 14 Angle of attack versus speed
for straight and level flight and for a 2-g turn.

Wing vortices
One might ask what the downwash from a wing looks like. The downwash comes off
the wing as a sheet and is related to the details on the load distribution on the wing.
Figure 15 shows, through condensation, the distribution of lift on an airplane during a
high-g maneuver. From the figure one can see that the distribution of load changes
from the root of the wing to the tip. Thus, the amount of air in the downwash must
also change along the wing. The wing near the root is "scooping" up much more air
than the tip. Since the root is diverting so much air the net effect is that the downwash
sheet will begin to curl outward around itself, just as the air bends around the top of
the wing because of the change in the velocity of the air. This is the wing vortex. The
tightness of the curling of the wing vortex is proportional to the rate of change in lift
along the wing. At the wing tip the lift must rapidly become zero causing the tightest
curl. This is the wing tip vortex and is just a small (though often most visible) part of
the wing vortex. Returning to figure 6 one can clearly see the development of the
wing vortices in the downwash as well as the wing tip vortices.

Fig 15 Condensation showing the distribution of lift along a wing.


The wingtip vortices are also seen.
(from Patterns in the Sky, J.F. Campbell and J.R. Chambers, NASA SP-514.)

Winglets (those small vertical extensions on the tips of some wings) are used to
improve the efficiency of the wing by increasing the effective length of the wing. The
lift of a normal wing must go to zero at the tip because the bottom and the top
communicate around the end. The winglets blocks this communication so the lift can
extend farther out on the wing. Since the efficiency of a wing increases with length,
this gives increased efficiency. One caveat is that winglet design is tricky and winglets
can actually be detrimental if not properly designed.
Ground effect
Another common phenomenon that is misunderstood is that of ground effect. That is
the increased efficiency of a wing when flying within a wing length of the ground. A
low-wing airplane will experience a reduction in drag by 50% just before it touches
down. There is a great deal of confusion about ground effect. Many pilots (and the
FAA VFR Exam-O-Gram No. 47) mistakenly believe that ground effect is the result
of air being compressed between the wing and the ground.

To understand ground effect it is necessary to have an understanding of upwash. For


the pressures involved in low speed flight, air is considered to be non-compressible.
When the air is accelerated over the top of the wing and down, it must be replaced. So
some air must shift around the wing (below and forward, and then up) to compensate,
similar to the flow of water around a canoe paddle when rowing. This is the cause of
upwash.

As stated earlier, upwash is accelerating air in the wrong direction for lift. Thus a
greater amount of downwash is necessary to compensate for the upwash as well as to
provide the necessary lift. Thus more work is done and more power required. Near the
ground the upwash is reduced because the ground inhibits the circulation of the air
under the wing. So less downwash is necessary to provide the lift. The angle of attack
is reduced and so is the induced power, making the wing more efficient.

Earlier, we estimated that a Cessna 172 flying at 110 knots must divert about 2.5
ton/sec to provide lift. In our calculations we neglected the upwash. From the
magnitude of ground effect, it is clear that the amount of air diverted is probably more
like 5 ton/sec.

Conclusions
Let us review what we have learned and get some idea of how the physical description
has given us a greater ability to understand flight. First what have we learned:

 The amount of air diverted by the wing is proportional to the speed of the wing
and the air density.
 The vertical velocity of the diverted air is proportional to the speed of the wing
and the angle of attack.
 The lift is proportional to the amount of air diverted times the vertical velocity
of the air.
 The power needed for lift is proportional to the lift times the vertical velocity of
the air.
Now let us look at some situations from the physical point of view and from the
perspective of the popular explanation.

 The plane’s speed is reduced. The physical view says that the amount of air
diverted is reduced so the angle of attack is increased to compensate. The
power needed for lift is also increased. The popular explanation cannot address
this.
 The load of the plane is increased. The physical view says that the amount of
air diverted is the same but the angle of attack must be increased to give
additional lift. The power needed for lift has also increased. Again, the popular
explanation cannot address this.
 A plane flies upside down. The physical view has no problem with this. The
plane adjusts the angle of attack of the inverted wing to give the desired lift.
The popular explanation implies that inverted flight is impossible.

As one can see, the popular explanation, which fixates on the shape of the wing, may
satisfy many but it does not give one the tools to really understand flight. The physical
description of lift is easy to understand and much more powerful.

Darrieus Vertical Axis Wind Turbine


A Darrieus is a type of vertical axis wind turbine (VAWT) generator. Unlike the Savonius
wind turbine, the Darrieus is a lift-type VAWT. Rather than collecting the wind in
cups dragging the turbine around, a Darrieus uses lift forces generated by the wind
hitting aerofoils to create rotation.
Benefits of the Darrieus Wind Turbine
A Darrieus wind turbine can spin at many times the speed of the wind hitting it (i.e.
the tip speed ratio(TSR) is greater than 1). Hence a Darrieus wind turbine generates
less torque than a Savonius but it rotates much faster. This makes Darrieus wind
turbines much better suited to electricity generation rather than water pumping and
similar activities. The centrifugal forces generated by a Darrieus turbine are very large
and act on the turbine blades which therefore have to be very strong – however the
forces on the bearings and generator are usually lower than are the case with a
Savonius.

Self Starting Darrieus Wind Turbines

Darrieus wind turbines are not self-starting. Therefore a small powered motor is
required to start off the rotation, and then when it has enough speed the wind passing
across the aerofoils starts to generate torque and the rotor is driven around by the wind.
An alternative is shown in the illustration above. Two small Savonius rotors are
mounted on the shaft of the Darrieus turbine to start rotation. These slow down the
Darrieus turbine when it gets going however they make the whole device a lot simpler
and easier to maintain.

Savonius Wind Turbines


A Savonius is a type of vertical axis wind turbine (VAWT) generator invented in 1922
by Sigurd Johannes Savonius from Finland though similar wind turbine designs had
been attempted in previous centuries.
The Savonius is a drag-type VAWT which operates in the same way as a cup
anemometer (pictured below).

Looking at the anemometer, it has three cups mounted on a rotor which is free to spin.
At any time the front of one of the cups will be facing into the wind, while the remaining
two cups will be back on to the wind. As the backs of each cup are rounded, they
experience much less drag than the front of the cup which is facing into the wind.
Therefore the force exerted by the wind on the open cup will be greater than the total
force exterted on the backs of the other cups and so the rotor will be pushed around
(anti-clockwise in the example of the pictured anemometer).

When the rotor has spun around one-third of a


revolution, the cup which was open to the wind will now be back on, and one of the
previously back on to the wind cups will have spun around on the rotor to offer its open
face into the wind. In this way, the rotor will continue to spin for as long as the wind is
blowing.
With a Savonius wind turbine it does not matter from which direction the wind is
blowing, since there will always be more force exerted on whichever cup has its open
face into the wind, and this will push the rotor around. This makes this design of wind
turbine ideal for areas with very turbulent wind.

Unfortunately, Savonius wind turbines typically only have an efficiency of around


15% – i.e. just 15% of the wind energy hitting the rotor is turned into rotational
mechanical energy. (This is much less than can be achieved with a Darrieus wind
turbine which uses lift rather than drag.)

The speed of the cups of a cup anemometer (and a Savonius wind turbine) cannot
rotate faster than the speed of the wind they are in and so they have a tip speed
ratio (TSR) of 1 or below. This means that a Savonius type vertical axis wind turbines
will turn slowly but generate high torque. Therefore Savonius turbines are not ideal
for electricity generation since turbine generators need to be turned at hundreds of
RPM to generate high voltages and currents. A gearbox can be employed to reduce the
torque and increase the RPM of the generator, but that leaves the Savonius requiring a
stronger wind to get spinning meaning it may not be able to self-start.

Savonius wind turbines are best suited to applications such as pumping water(see
image above) and grinding grain for which slow rotation and high torque are desireable.
They are often used in the Third World in exactly these ways with wind turbines made
from an old oil barrel (or barrels) welded to a pole which is passed through bearings
taken from an old vehicle.

Because of the high torque yield of a Savonius wind turbine, the bearings used must
be very sturdy and may require servicing every couple of years.
Savonius Wind Turbines for Electricity Generation

Though not ideal, Savonius wind turbines can and are used in electricity
generation – for example see our article on the Envirotek V20 VAWT generator – with
the benefit that they continue to generate electricity in the strongest winds without being
damaged, they are very quiet, and they are relatively easy to make.

Standard HAWT (windmill) type wind turbines are unsuitable in locations with strong
turbulent winds, so Savonius wind turbines can sometimes be the best option. They can
also be mounted relatively safely at ground level as they spin much much slower than
the speed of the tips of an HAWT.

Make a Savonius Wind Turbine


Savonius turbines do not scale well to kW sizes, but for a small project they are typically
the easiest and cheapest wind turbine generators to build yourself. Click here to view
our article Build a Savonius Wind Turbine to find out how to put together a small
(<100W) Savonius wind turbine using cheap and easy to find items, and only a few
tools.

Airplane Wings
https://www.grc.nasa.gov/www/k-12/airplane/right2.html
The position of the trailing edge flaps on a typical airliner. In this picture, the flaps are extended, note also the
drooped leading edge slats.

The three orange pods are fairings streamlining the flap track mechanisms. The flaps (two on each side, on the
Airbus A319) lie directly above these.
British Airways Boeing 757-200 landing. The flap track A close look at the spoiler (the parts of the wing that are
fairings are the canoe-shaped fairings that protect and raised up) during the landing of an Airbus A321.
streamline the flap operating mechanisms

<< In aeronautics, a spoiler (sometimes called a lift spoiler or lift dumper) is a device intended to
intentionally reduce the lift component of an airfoil in a controlled way. Most often, spoilers are plates on
the top surface of a wing that can be extended upward into the airflow to spoil it. By so doing, the spoiler
creates a controlled stall over the portion of the wing behind it, greatly reducing the lift of that wing
section. Spoilers differ from airbrakes in that airbrakes are designed to increase drag without affecting lift,
while spoilers reduce lift as well as increasing drag.>>
This page shows the parts of an airplane and their functions. Airplanes are
transportation devices which are designed to move people and cargo from one place to
another. Airplanes come in many different shapes and sizes depending on the mission
of the aircraft. The airplane shown on this slide is a turbine-powered airliner which has
been chosen as a representative aircraft.

For any airplane to fly, one must lift the weight of the airplane itself, the fuel, the
passengers, and the cargo. The wings generate most of the lift to hold the plane in the
air. To generate lift, the airplane must be pushed through the air. The air resists the
motion in the form of aerodynamic drag. Modern airliners use winglets on the tips of the
wings to reduce drag. The turbine engines, which are located beneath the wings,
provide the thrust to overcome drag and push the airplane forward through the air.
Smaller, low-speed airplanes use propellers for the propulsion system instead of turbine
engines.

To control and maneuver the aircraft, smaller wings are located at the tail of the plane.
The tail usually has a fixed horizontal piece, called the horizontal stabilizer, and a fixed
vertical piece, called the vertical stabilizer. The stabilizers' job is to provide stability for
the aircraft, to keep it flying straight. The vertical stabilizer keeps the nose of the plane
from swinging from side to side, which is called yaw. The horizontal stabilizer prevents
an up-and-down motion of the nose, which is called pitch. (On the Wright brother's first
aircraft, the horizontal stabilizer was placed in front of the wings. Such a configuration is
called a canard after the French word for "duck").

At the rear of the wings and stabilizers are small moving sections that are attached to
the fixed sections by hinges. In the figure, these moving sections are colored
brown. Changing the rear portion of a wing will change the amount of force that the wing
produces. The ability to change forces gives us a means of controlling and maneuvering
the airplane. The hinged part of the vertical stabilizer is called the rudder; it is used to
deflect the tail to the left and right as viewed from the front of the fuselage. The hinged
part of the horizontal stabilizer is called the elevator; it is used to deflect the tail up and
down. The outboard hinged part of the wing is called the aileron; it is used to roll the
wings from side to side. Most airliners can also be rolled from side to side by using
the spoilers. Spoilers are small plates that are used to disrupt the flow over the wing and
to change the amount of force by decreasing the lift when the spoiler is deployed.

The wings have additional hinged, rear sections near the body that are
called flaps. Flaps are deployed downward on takeoff and landing to increase the
amount of force produced by the wing. On some aircraft, the front part of the wing will
also deflect. Slats are used at takeoff and landing to produce additional force.
The spoilers are also used during landing to slow the plane down and to counteract the
flaps when the aircraft is on the ground. The next time you fly on an airplane, notice how
the wing shape changes during takeoff and landing.
The fuselage or body of the airplane, holds all the pieces together. The pilots sit in
the cockpit at the front of the fuselage. Passengers and cargo are carried in the rear of
the fuselage. Some aircraft carry fuel in the fuselage; others carry the fuel in the wings.

As mentioned above, the aircraft configuration in the figure was chosen only as an
example. Individual aircraft may be configured quite differently from this airliner. The
Wright Brothers 1903 Flyer had pusher propellers and the elevators at the front of the
aircraft. Fighter aircraft often have the jet engines buried inside the fuselage instead of
in pods hung beneath the wings. Many fighter aircraft also combine the horizontal
stabilizer and elevator into a single stabilator surface. There are many possible aircraft
configurations, but any configuration must provide for the four forces needed for flight.
A force may be thought of as a push or pull in a specific direction. A force is a vector
quantity so a force has both a magnitude and a direction. When describing forces, we
have to specify both the magnitude and the direction. This slide shows the forces that
act on an airplane in flight.

Weight
Weight is a force that is always directed toward the center of the earth.
The magnitude of the weight depends on the mass of all the airplane parts, plus the
amount of fuel, plus any payload on board (people, baggage, freight, etc.). The weight is
distributed throughout the airplane. But we can often think of it as collected and acting
through a single point called the center of gravity. In flight, the airplane rotates about
the center of gravity.

Flying encompasses two major problems; overcoming the weight of an object by some
opposing force, and controlling the object in flight. Both of these problems are related to
the object's weight and the location of the center of gravity. During a flight, an
airplane's weight constantly changes as the aircraft consumes fuel. The distribution of
the weight and the center of gravity also changes. So the pilot must constantly adjust
the controls to keep the airplane balanced, or trimmed.

Lift
To overcome the weight force, airplanes generate an opposing force called lift. Lift is
generated by the motion of the airplane through the air and is an aerodynamic
force. "Aero" stands for the air, and "dynamic" denotes motion. Lift is
directed perpendicular to the flight direction. The magnitude of the lift depends on
several factors including the shape, size, and velocity of the aircraft. As with weight,
each part of the aircraft contributes to the aircraft lift force. Most of the lift is generated
by the wings. Aircraft lift acts through a single point called the center of pressure. The
center of pressure is defined just like the center of gravity, but using
the pressure distribution around the body instead of the weight distribution.

The distribution of lift around the aircraft is important for solving the control problem.
Aerodynamic surfaces are used to control the aircraft in roll, pitch, and yaw.

Drag
As the airplane moves through the air, there is another aerodynamic force present. The
air resists the motion of the aircraft and the resistance force is called drag. Drag is
directed along and opposed to the flight direction. Like lift, there are manyfactors that
affect the magnitude of the drag force including the shape of the aircraft,
the "stickiness" of the air, and thevelocity of the aircraft. Like lift, we collect all of the
individual components' drags and combine them into a single aircraft drag magnitude.
And like lift, drag acts through the aircraft center of pressure.

Thrust
To overcome drag, airplanes use a propulsion system to generate a force
called thrust. The direction of the thrust force depends on how the engines are attached
to the aircraft. In the figure shown above, two turbine engines are located under the
wings, parallel to the body, with thrust acting along the body centerline. On some
aircraft, such as the Harrier, the thrust direction can be varied to help the airplane take
off in a very short distance. The magnitude of the thrust depends on many factors
associated with the propulsion system including the type of engine, the number of
engines, and the throttle setting.

Lift is the force that directly opposes the weight of an airplane and holds the airplane in
the air. Lift is generated by every part of the airplane, but most of the lift on a normal
airliner is generated by the wings. Lift is a mechanical aerodynamic force produced by
the motion of the airplane through the air. Because lift is a force, it is a vector quantity,
having both a magnitude and a direction associated with it. Lift acts through the center
of pressure of the object and is directed perpendicular to the flow direction. There are
several factors which affect the magnitude of lift.

HOW IS LIFT GENERATED?

There are many explanations for the generation of lift found in encyclopedias, in basic
physics textbooks, and on Web sites. Unfortunately, many of the explanations are
misleading and incorrect. Theories on the generation of lift have become a source of
great controversy and a topic for heated arguments. To help you understand lift and its
origins, a series of pages will describe the various theories and how some of the
popular theories fail.

Lift occurs when a moving flow of gas is turned by a solid object. The flow is turned in
one direction, and the lift is generated in the opposite direction, according to Newton's
Third Law of action and reaction. Because air is a gas and the molecules are free to
move about, any solid surface can deflect a flow. For an aircraft wing, both the upper
and lower surfaces contribute to the flow turning. Neglecting the upper surface's part in
turning the flow leads to an incorrect theory of lift.

NO FLUID, NO LIFT

Lift is a mechanical force. It is generated by the interaction and contact of a solid body
with a fluid (liquid or gas). It is not generated by a force field, in the sense of
a gravitational field,or an electromagnetic field, where one object can affect another
object without being in physical contact. For lift to be generated, the solid body must be
in contact with the fluid: no fluid, no lift. The Space Shuttle does not stay in space
because of lift from its wings but because of orbital mechanics related to its speed.
Space is nearly a vacuum. Without air, there is no lift generated by the wings.

NO MOTION, NO LIFT

Lift is generated by the difference in velocity between the solid object and the fluid.
There must be motion between the object and the fluid: no motion, no lift. It makes no
difference whether the object moves through a static fluid, or the fluid moves past a
static solid object. Lift acts perpendicular to the motion. Drag acts in the direction
opposed to the motion.

When two solid objects interact in a mechanical process, forces are transmitted, or
applied, at the point of contact. But when a solid object interacts with a fluid, things are
more difficult to describe because the fluid can change its shape. For a solid body
immersed in a fluid, the "point of contact" is every point on the surface of the body. The
fluid can flow around the body and maintain physical contact at all points. The
transmission, or application, of mechanical forces between a solid body and a fluid
occurs at every point on the surface of the body. And the transmission occurs through
the fluid pressure.

Variation in Pressure

The magnitude of the force acting over a small section of an object immersed in a fluid
equals the pressure p times the area A of the section. A quick units check shows that:

p * A = (force/area) * area = force

As discussed on the fluid pressure slide, pressure is a scalar quantity related to the
momentum of the molecules of a fluid. Since a force is a vector quantity, having both
magnitude and direction, we must determine the direction of the pressure force.
Pressure acts perpendicular (or normal) to the solid surface of an object. So the
direction of the force on the small section of the object is along the normal to the
surface. We denote this direction by the letter n.

The normal direction changes from the front of the airfoil to the rear and from the top to
the bottom. We indicate this variation on the figure by several small arrows pointing
perpendicular to the surface and labeled with an n. To obtain the net mechanical force
over the entire solid object, we must sum the contributions from all the small sections.
Mathematically, the summation is indicated by the Greek letter sigma ( ) The net
aerodynamic force F is equal to the sum of the product of the pressure p times the
incremental area delta A in the normal direction n.

F= p * n * delta A

In the limit of infinitely small sections, this gives the integral of the pressure times the
area around the closed surface. Using the symbol S dA for integration, we have:

F = S (p * n) dA

where the integral is taken all around the body. On the figure, that is why the integral
sign has a circle through it.

If the pressure on a closed surface is a constant, there is no net force produced


because the summation of the directions of the normal adds up to zero. For every small
section there is another small section whose normal points in exactly the opposite
direction.

F = S (p * n) dA = p * S n dA = 0
For a fluid in motion, the velocity has different values at different locations around the
body. The local pressure is related to the local velocity, so the pressure also varies
around the closed surface and a net force is produced. On the figure at the lower right,
we show the variation of the pressure around the airfoil as obtained by a solution of
the Euler equations. The blue line shows the variation from front to back on the lower
surface, while the red line shows the variation from front to back on the upper surface,
The black line gives the reference free stream pressure. Summing the pressure
perpendicular to the surface times the area around the body produces a net force.

F = S (p * n) dA

Definitions of Lift and Drag

Since the fluid is in motion, we can define a flow direction along the motion.
The component of the net force perpendicular (or normal) to the flow direction is called
the lift; the component of the net force along the flow direction is called the drag. These
are definitions. In reality, there is a single, net, integrated force caused by the pressure
variations along a body. This aerodynamic force acts through the average location of
the pressure variation which is called the center of pressure.

Velocity Distribution

For an ideal fluid with no boundary layers, the surface of an object is a streamline. If the
velocity is low, and no energy is added to the flow, we can use Bernoulli's
equation along a streamline to determine the pressure distribution for a known velocity
distribution. If boundary layers are present, things are a little more confusing, since the
external flow responds to the edge of the boundary layer and the pressure on the
surface is imposed from the edge of the boundary layer. If the boundary layer separates
from the surface, it gets even more confusing. How do we determine the velocity
distribution around a body? Specifying the velocity is the source of error in two of the
more popular incorrect theories of lift. To correctly determine the velocity distribution, we
have to solve equations expressing a conservation of mass, momentum, and energy for
the fluid passing the object. In some cases, we can solve simplified versions of the
equations to determine the velocity and pressure.

Summary

To summarize, for any object immersed in a fluid, the mechanical forces are transmitted
at every point on the surface of the body. The forces are transmitted through the
pressure, which acts perpendicular to the surface. The net force can be found by
integrating (or summing) the pressure times the area around the entire surface. For a
moving flow, the pressure will vary from point to point because the velocity varies from
point to point. For some simple flow problems, we can determine the pressure
distribution (and the net force) if we know the velocity distribution by using Bernoulli's
equation.
Lift can be generated by a wide variety of objects, including airplane wings, rotating
cylinders, spinning balls, and flat plates. Lift is the force that holds an aircraft in the air.
Lift can be generated by any part of the airplane, but most of the lift on a normal airliner
is generated by the wings. How is lift generated?

Force = Mass x Acceleration

Lift is a force. From Newton's second law of motion, a force F is produced when a
mass m is accelerated a:

F=m*a

An acceleration is a change in velocity V with a change in time t.

F = m * (V1 - V0) / (t1 - t0)

We have written this relationship as a difference equation, but it is recognized that the
relation is actually a differential from calculus.

F = m * dV/dt

The important fact is that a force causes a change in velocity; and, likewise, a change in
velocity generates a force. The equation works both ways. A velocity has both
a magnitude called the speed and a direction associated with it. Scientists and
mathematicians call this a vector quantity. So, to change either the speed or the
direction of a flow, you must impose a force. And if either the speed or the direction of a
flow is changed, a force is generated.

Lift Generated in a Moving Fluid

For a body immersed in a moving fluid, the fluid remains in contact with the surface of
the body. If the body is shaped, moved, or inclined in such a way as to produce a net
deflection or turning of the flow, the local velocity is changed in magnitude, direction, or
both. Changing the velocity creates a net force on the body. It is very important to note
that the turning of the fluid occurs because the molecules of the fluid stay in contact with
the solid body since the molecules are free to move. Any part of the solid body can
deflect a flow. Parts facing the oncoming flow are said to be windward, and parts facing
away from the flow are said to be leeward. Both windward and leeward parts deflect a
flow. Ignoring the leeward deflection leads to a popular incorrect theory of lift.

Changes in Speed or Direction

Lift is a force generated by turning a flow. Since a force is a vector quantity (like the
velocity), it has both a magnitude and a direction. The direction of the lift force is
defined to be perpendicular to the initial flow direction. (The drag is defined to be
along the flow direction.) The magnitude depends on several factors concerning the
object and the flow.

Summary

Lift and drag are mechanical forces generated on the surface of an object as it interacts
with a fluid. The net fluid force is generated by the pressure acting over the entire
surface of a closed body. The pressure varies around a body in a moving fluid because
it is related to the fluid momentum (mass times velocity). The velocity varies around
the body because of the flow deflection described above.
Since we live in a three dimensional world, it is necessary to control the attitude or
orientation of a flying aircraft in all three dimensions. In flight, any aircraft will rotate
about its center of gravity, a point which is the average location of the mass of the
aircraft. We can define a three dimensional coordinate system through the center of
gravity with each axis of this coordinate system perpendicular to the other two axes. We
can then define the orientation of the aircraft by the amount of rotation of the parts of the
aircraft along these principal axes.

The yaw axis is defined to be perpendicular to the plane of the wings with its origin at
the center of gravity and directed towards the bottom of the aircraft. A yaw motion is a
movement of the nose of the aircraft from side to side. The pitch axis is perpendicular
to the yaw axis and is parallel to the plane of the wings with its origin at the center of
gravity and directed towards the right wing tip. A pitch motion is an up or down
movement of the nose of the aircraft. The roll axis is perpendicular to the other two
axes with its origin at the center of gravity, and is directed towards the nose of the
aircraft. A rolling motion is an up and down movement of the wing tips of the aircraft.

In flight, the control surfaces of an aircraft produce aerodynamic forces. These forces
are applied at the center of pressure of the control surfaces which are some distance
from the aircraft cg and produce torques (or moments) about the principal axes. The
torques cause the aircraft to rotate. The elevators produce a pitching moment,
the rudder produces a yawing moment, and the ailerons produce a rolling moment. The
ability to vary the amount of the force and the moment allows the pilot to maneuver or
to trim the aircraft. The first aircraft to demonstrate active control about all three axes
was the Wright brothers' 1902 glider.

All that is necessary to create lift is to turn a flow of air. An aerodynamic, curved airfoil
will turn a flow. But so will a simple flat plate, if it is inclined to the flow. The fuselage of
an airplane will also generate lift if it is inclined to the flow. For that matter, an
automobile body also turns the flow through which it moves, generating a lift force. Lift is
a big problem for NASCAR racing machines and race cars now include spoilers on the
roof to kill lift in a spin. Any physical body moving through a fluid can create lift if it
produces a net turning of the flow.

There are many factors that affect the turning of the flow, which creates lift. We can
group these factors into(a) those associated with the object, (b) those associated with
the motion of the object through the air, and (c) those associated with the air itself:

1. Object: At the top of the figure, aircraft wing geometry has a large effect on the
amount of lift generated. The airfoil shape and wing size will both affect the
amount of lift. The ratio of the wing span to the wing area also affects the amount
of lift generated by a wing.
2. Motion: To generate lift, we have to move the object through the air. The lift then
depends on the velocity of the air and how the object is inclined to the flow.
3. Air: Lift depends on the mass of the flow. The lift also depends in a complex way
on two other properties of the air: its viscosity and its compressibility.

We can gather all of this information on the factors that affect lift into a single
mathematical equation called the Lift Equation.With the lift equation we can predict how
much lift force will be generated by a given body moving at a given speed.

Lift is created by deflecting a flow of air and drag is generated on a body in a wide
variety of ways. From Newton's second law of motion, the aerodynamic forces on the
body (lift and drag) are directly related to the change in momentum of the fluid with time.
The fluid momentum is equal to the mass times the velocity of the fluid. Since the air
moves, defining the mass gets a little tricky and aerodynamicists usually relate the
effect of mass on lift and drag to the air density. The mathematical derivation for this
conversion is given on another slide dealing with momentum effects on lift. As a result
of this derivation, we find that lift and drag depend on the square of the velocity.

The velocity used in the lift and drag equations is the relative velocity between an object
and the flow. Since the aerodynamic force depends on the square of the velocity,
doubling the velocity will quadruple the lift and drag.

Lift is created by deflecting a moving fluid (liquid or gas), and drag is generated on a
body in a wide variety of ways. From Newton's second law of motion, the aerodynamic
forces on the body (lift and drag) are directly related to the change in momentum of the
fluid with time. The fluid momentum is equal to the mass times the velocity of the fluid.
Since the fluid is moving, defining the mass gets a little tricky. If the mass of fluid were
brought to a halt, it would occupy some volume in space; and we could define
its density to be the mass divided by the volume. With a little math which is described
on the fluid momentum page, we can show that the aerodynamic forces are directly
proportional to the density of the fluid that flows by the airfoil.

Lift and drag depend linearly on the density of the fluid. Halving the density halves the
lift, halving the density halves the drag. The fluid density depends on the type of fluid
and the depth of the fluid. In the atmosphere, air density decreases as altitude
increases. This explains why airplanes have a flight ceiling, an altitude above which it
cannot fly. As an airplane ascends, a point is eventually reached where there just isn't
enough air mass to generate enough lift to overcome the airplane's weight.
The relation between altitude and density is a fairly complex exponential that has been
determined by measurements in the atmosphere.

The amount of lift generated by an object depends on how much the flow is turned,
which depends on the shape of the object. In general, the lift is a very complex function
of the shape. Aerodynamicists model the shape effect by a lift coefficientwhich is
normally determined through wind tunnel testing. For some simple shapes, we can
develop mathematical equationsto determine the lift coefficient. The simplest model, the
two dimensional Kutta-Joukowski airfoil, is studied by undergraduate students.
The FoilSim computer program provides the results of this analysis in a form readily
usable by students. A result of the analysis shows that the greater the flow turning,
the greater the lift generated by an airfoil.

This slide shows the flow fields for two different airfoils. The airfoil on the left is a
symmetric airfoil; the shapes above and below the white centerline are the same. The
airfoil on the right is curved near the trailing edge. The yellow lines on each figure show
the streamlines of flow from left to right. The left figure shows no net turning of the flow
and produces no lift; the right figure shows a large amount of turning and generates a
large amount of lift. The front portions of both airfoils are nearly identical. The aft portion
of the right airfoil creates the higher turning.

The example shown above explains why the aft portion of wings have hinged
sections to control and maneuver an aircraft. Deflecting the aft section down produces a
geometry similar to the figure on the right producing more lift. Similarly, if the aft section
is deflected up, it creates less lift (or even negative lift). The ability to vary the amount of
lift over a portion of the wing gives the pilot the ability to maneuver an aircraft. The
following slides show the deflection of the control surfaces and the resulting motion of
the aircraft:

 Elevator controls pitching motion.


 Rudder controls yawing motion.
 Ailerons control rolling motion.

The amount of lift generated by an object depends on the size of the object. Lift is
an aerodynamic force and therefore depends on the pressure variation of the air
around the body as it moves through the air. The total aerodynamic force is equal to the
pressure times the surface area around the body. Lift is the component of this force
perpendicular to the flight direction. Like the other aerodynamic force, drag, the lift is
directly proportional to the area of the object. Doubling the area doubles the lift.

There are several different areas from which to choose when developing the reference
area used in the lift equation. Since most of the lift is generated by the wings, and lift is
the force perpendicular to the flight direction, the logical choice is the wing planform
area. The planform area is the area of the wing as viewed from above the wing, looking
along the "lift" direction. It is a flat plane, and is NOT the total surface area (top and
bottom) of the entire wing, although it is almost half that number for most wings. We
could, in theory, use the total surface area as the reference area. The total surface area
is proportional to the wing planform area. Since the lift coefficient is determined
experimentally, by measuring the lift and measuring the area and performing the
necessary math to produce the coefficient, we are free to use any area which can be
easily measured. If we choose the total surface area, the computed coefficient has a
different value than if we choose the wing planform area, but the lift is the same, and the
coefficients are related by the ratio of the areas.

This slide shows the projected surface area for two different aircraft. The airplane on the
left is shown in a cruise condition while the airplane on the right is shown in a takeoff or
landing condition. Takeoff and landing are times of relatively low velocity, so to keep the
lift high (to avoid the ground!) designers try to increase the wing area. This is done by
sliding the flaps backwards along metal tracks and shifting the slats forward to increase
the wing area. The next time you fly in an airliner, watch the wings during takeoff and
landing to see the change in wing area.

Since aircraft come in many shapes and sizes, an aerodynamicist has to be able to
compute the wing area for many different shapes. Most of these skills are learned in
high school.

Lift depends on the density of the air, the square of the velocity, the air's viscosity and
compressibility, the surface area over which the air flows, the shape of the body, and
the body's inclination to the flow. In general, the dependence on body shape, inclination,
air viscosity, and compressibility is very complex.
One way to deal with complex dependencies is to characterize the dependence by a
single variable. For lift, this variable is called the lift coefficient, designated "Cl." This
allows us to collect all the effects, simple and complex, into a single equation. The lift
equation states that lift L is equal to the lift coefficient Cl times the density r times half of
the velocity V squared times the wing area A.

L = Cl * A * .5 * r * V^2

For given air conditions, shape, and inclination of the object, we have to determine a
value for Cl to determine the lift. For some simple flow conditions and geometries and
low inclinations, aerodynamicists can determine the value of Cl mathematically. But, in
general, this parameter is determined experimentally.

In the equation given above, the density is designated by the letter "r." We do not use
"d" for density, since "d" is often used to specify distance. In many textbooks on
aerodynamics, the density is given by the Greek symbol "rho" (Greek for "r"). The
combination of terms "density times the square of the velocity divided by two" is called
the dynamic pressure and appears in Bernoulli's pressure equation.

The lift coefficient is a number that aerodynamicists use to model all of the complex
dependencies of shape, inclination, and some flow conditions on lift. This equation is
simply a rearrangement of the lift equation where we solve for the lift coefficient in terms
of the other variables. The lift coefficient Cl is equal to the lift L divided by the quantity:
density r times half the velocity V squared times the wing area A.

Cl = L / (A * .5 * r * V^2)

The quantity one half the density times the velocity squared is called the dynamic
pressure q. So

Cl = L / (q * A)

The lift coefficient then expresses the ratio of the lift force to the force produced by the
dynamic pressure times the area.

Here is a way to determine a value for the lift coefficient. In a controlled


environment (wind tunnel) we can set the velocity, density, and area and measure the
lift produced. Through division, we arrive at a value for the lift coefficient. We can then
predict the lift that will be produced under a different set of velocity, density
(altitude), and area conditions using the lift equation.

The lift coefficient contains the complex dependencies of object shape on lift. For three
dimensional wings, the downwash generated near the wing tips reduces the overall lift
coefficient of the wing. The lift coefficient also contains the effects of air viscosity and
compressibility. To correctly use the lift coefficient, we must be sure that the viscosity
and compressibility effects are the same between our measured case and the predicted
case. Otherwise, the prediction will be inaccurate.

For very low speeds (< 200 mph) the compressibility effects are negligible. At higher
speeds, it becomes important to match Mach numbers between the two cases. Mach
number is the ratio of the velocity to the speed of sound. So it is completely incorrect to
measure a lift coefficient at some low speed (say 200 mph) and apply that lift coefficient
at twice the speed of sound (approximately 1,400 mph, Mach = 2.0). The compressibility
of the air will alter the important physics between these two cases.

Similarly, we must match air viscosity effects, which becomes very difficult. The
important matching parameter for viscosity is the Reynolds number. The Reynolds
number expresses the ratio of inertial forces to viscous forces. If the Reynolds number
of the experiment and flight are close, then we properly model the effects of the viscous
forces relative to the inertial forces. If they are very different, we do not correctly model
the physics of the real problem and will predict an incorrect lift.
Lift is created by deflecting a flow of air, and drag is generated on a body in a wide
variety of ways. From Newton's second law of motion, the aerodynamic force F on the
body is directly related to the change in momentum of the fluid with time t. The fluid
momentum is equal to the mass m times the velocity V of the fluid.

F = d (m * V) / dt

F = constant * V * m / t

Since the air moves, defining the mass is tricky. If the mass of fluid were brought to a
halt, it would occupy some volume in space. We can define the density (r) of the fluid to
be the mass divided by the volume v.

r=m/v

Since the fluid is moving, we must determine the mass in terms of the mass flow rate.
The mass flow rate is the amount of mass passing a given point during some time
interval t and its units are mass/time. We can relate the mass flow rate to the density
mathematically. The mass flow rate mdot is equal to the density times the velocity times
the area A through which the mass passes.

mdot = m / t = r * V * A

With knowledge of the mass flow rate, we can express the aerodynamic force as equal
to the mass flow rate times the velocity.

F = constant * V * r * V * A

A quick units check:

mass * length / time^2 = constant * length/time * mass/length^3 * length/time * length^2


mass * length / time^2 = mass * length/time^2

Combining the velocity dependence and absorbing the area into the constant, we find:

F = constant * r * V^2

The aerodynamic force equals a constant times the density times the velocity squared.
The dynamic pressure of a moving flow is equal to one half of the density times the
velocity squared. The aerodynamic force is directly proportional to the dynamic pressure
of the flow.

Effect of Velocity on Aerodynamic Forces

The velocity used in the aerodynamic equation is the relative velocity between an object
and the flow. The aerodynamic force depends on the square of the velocity. Doubling
the velocity quadruples the force. The dependence of lift and drag on the square of
the velocity has been known for more than a hundred years. The Wright brothers used
this information in the design of their first aircraft.

Effect of Air Density on Aerodynamic Forces

The aerodynamic force depends linearly on the density of the air. Halving the density
halves the force. As altitude increases, the air density decreases. This explains why
airplanes have a flight ceiling, an altitude above which it cannot fly. As an airplane
ascends, a point is reached where there is not enough air mass to generate enough lift
to overcome the airplane's weight. The relation between altitude and density is a fairly
complex exponential.
Drag depends on the density of the air, the square of the velocity, the air's viscosity and compressibility,
the size and shape of the body, and the body's inclination to the flow. In general, the dependence on body
shape, inclination, air viscosity, and compressibility is very complex.

One way to deal with complex dependencies is to characterize the dependence by a single variable. For
drag, this variable is called the drag coefficient, designated "Cd." This allows us to collect all the effects,
simple and complex, into a single equation. The drag equation states that drag D is equal to the drag
coefficient Cd times the density r times half of the velocity V squared times the reference area A.

D = Cd * A * .5 * r * V^2

For given air conditions, shape, and inclination of the object, we must determine a value for Cd to
determine drag. Determining the value of the drag coefficient is more difficult than determining the lift
coefficient because of the multiple sources of drag. The drag coefficient given above includes form drag,
skin friction drag, wave drag, and induced drag components. Ram drag is usually included in the net
thrust because it depends on the airflow through the engine. Drag coefficients are almost always
determined experimentally using a wind tunnel.

Notice that the area (A) given in the drag equation is given as a reference area. The drag depends
directly on the size of the body. Since we are dealing with aerodynamic forces, the dependence can be
characterized by some area. But which area do we choose? If we think of drag as being caused by
friction between the air and the body, a logical choice would be the total surface area of the body. If we
think of drag as being a resistance to the flow, a more logical choice would be the frontal area of the body
that is perpendicular to the flow direction. And finally, if we want to compare with the lift coefficient, we
should use the same wing area used to derive the lift coefficient. Since the drag coefficient is usually
determined experimentally by measuring drag and the area and then performing the division to produce
the coefficient, we are free to use any area that can be easily measured. If we choose the wing area,
rather than the cross-sectional area, the computed coefficient will have a different value. But the drag is
the same, and the coefficients are related by the ratio of the areas. In practice, drag coefficients are
reported based on a wide variety of object areas. In the report, the aerodynamicist must specify the area
used; when using the data, the reader may have to convert the drag coefficient using the ratio of the
areas.

In the equation given above, the density is designated by the letter "r." We do not use "d" for density
since "d" is often used to specify distance. In many textbooks on aerodynamics, density is given by the
Greek symbol "rho" (Greek for "r"). The combination of terms "density times the square of the velocity
divided by two" is called the dynamic pressure and appears in Bernoulli's pressure equation.

Lift depends on the density of the air, the square of the velocity, the air's viscosity and compressibility, the
surface area over which the air flows, the shape of the body, and the body's inclination to the flow. In
general, the dependence on body shape, inclination, air viscosity, and compressibility is very complex.

One way to deal with complex dependencies is to characterize the dependence by a single variable. For
lift, this variable is called the lift coefficient, designated "Cl." This allows us to collect all the effects, simple
and complex, into a single equation. The lift equation states that lift L is equal to the lift coefficient Cl
times the density r times half of the velocity V squared times the wing area A.

L = Cl * A * .5 * r * V^2

For given air conditions, shape, and inclination of the object, we have to determine a value for Cl to
determine the lift. For some simple flow conditions and geometries and low inclinations, aerodynamicists
can determine the value of Cl mathematically. But, in general, this parameter is determined
experimentally.
In the equation given above, the density is designated by the letter "r." We do not use "d" for density,
since "d" is often used to specify distance. In many textbooks on aerodynamics, the density is given by
the Greek symbol "rho" (Greek for "r"). The combination of terms "density times the square of the velocity
divided by two" is called the dynamic pressure and appears in Bernoulli's pressure equation.

Savonius wind turbine


From Wikipedia, the free encyclopedia

Savonius wind turbine

Savonius wind turbines are a type of vertical-axis wind turbine (VAWT), used for converting the
force of the wind into torque on a rotating shaft. The turbine consists of a number of aerofoils,
usually—but not always—vertically mounted on a rotating shaft or framework, either ground
stationed or tethered in airborne systems.

Origin
The Savonius wind turbine was invented by the Finnish engineer Sigurd Johannes Savonius in
1922. However, Europeans had been experimenting with curved blades on vertical wind turbines for
many decades before this. The earliest mention is by the Italian Bishop of Czanad, Fausto Veranzio,
who was also an engineer. He wrote in his 1616 book Machinae novae about several vertical axis
wind turbines with curved or V-shaped blades. None of his or any other earlier examples reached
the state of development made by Savonius. In his Finnish biography there is mention of his
intention to develop a turbine-type similar to the Flettner-type, but autorotationary. He experimented
with his rotor on small rowing vessels on lakes in his country. No results of his particular
investigations are known, but the Magnus effect is confirmed by König.[1] The two Savonius
patents: US1697574, were filed in 1925 by Sigurd Johannes Savonius, and US1766765, in 1928.

Operation
Schematic drawing of a two-scoop Savonius turbine

The Savonius turbine is one of the simplest turbines. Aerodynamically, it is a drag-type device,
consisting of two or three scoops. Looking down on the rotor from above, a two-scoop machine
would look like an "S" shape in cross section. Because of the curvature, the scoops experience less
drag when moving against the wind than when moving with the wind. The differential drag causes
the Savonius turbine to spin. Because they are drag-type devices, Savonius turbines extract much
less of the wind's power than other similarly-sized lift-type turbines. Much of the swept area of a
Savonius rotor may be near the ground, if it has a small mount without an extended post, making the
overall energy extraction less effective due to the lower wind speeds found at lower heights.

Power and rotational speed

According to Betz's law, the maximum power that is possible to extract from a rotor is

, where is the density of air, and are the height and radius of the rotor

and is the wind speed. However, in practice the extractable power is about half that [2] (one can
argue that only one half of the rotor — the scoop co-moving with the wind — works at each instant of

time). Thus, one gets .

The angular frequency of a rotor is given by , where is a dimensionless factor called


the tip-speed ratio. λ is a characteristic of each specific windmill, and for a Savonius rotor λ is
typically around unity.
For example, an oil-barrel sized Savonius rotor with h=1 m and r=0.5 m under a wind of v=10 m/s,
will generate a maximum power of 180 W and an angular speed of 20 rad/s (190 revolutions per
minute).

Use
Combined Darrieus-Savonius generator in Taiwan

Savonius turbines are used whenever cost or reliability is much more important than efficiency.
Most anemometers are Savonius turbines for this reason, as efficiency is irrelevant to the application
of measuring wind speed. Much larger Savonius turbines have been used to generate electric power
on deep-water buoys, which need small amounts of power and get very little maintenance. Design is
simplified because, unlike with horizontal axis wind turbines (HAWTs), no pointing mechanism is
required to allow for shifting wind direction and the turbine is self-starting. Savonius and other
vertical-axis machines are good at pumping water and other high torque, low rpm applications and
are not usually connected to electric power grids. In the early 1980s Risto Joutsiniemi developed
a helical rotor (wiki:fi) version that does not require end plates, has a smoother torque profile and is
self-starting in the same way a crossed pair of straight rotors is.
The most ubiquitous application of the Savonius wind turbine is the Flettner Ventilator, which is
commonly seen on the roofs of vans and buses and is used as a cooling device. The ventilator was
developed by the German aircraft engineer Anton Flettner in the 1920s. It uses the Savonius wind
turbine to drive an extractor fan. The vents are still manufactured in the UK by Flettner Ventilator
Limited.[3]
Small Savonius wind turbines are sometimes seen used as advertising signs where the rotation
helps to draw attention to the item advertised. They sometimes feature a simple two-
frame animation.

http://airfoiltools.com/search/airfoils

You might also like