Sei sulla pagina 1di 12

En el centro de presiones (centroide del airfoil), por definición, no hay momento

aunque no es fijo ya que varía según el angulo de ataque. El momento luego se


podrá obtener por sustitucipon en el centro aerodinámico tralado la fuerza y utilizo
el brazo de palanca del desplazamiento.

Equations (2.9a) to (2.9c) indicate that CL, CD, and for a body shape vary with the
angle of attack, Reynolds and Mach number.

Note: The lift coefficient at the much higher cruise velocity is much smaller than
that at takeoff, even though the density at 30,000 ft is smaller than that at sea level.
It is sometimes convenient to think that the lift at high speeds is mainly obtained
from the high dynamic pressure; hence only a small lift coefficient is required. In
turn, at low speeds the dynamic pressure is lower, and in order to keep the lift
equal to the weight in steady, level flight, the low dynamic pressure must be
compensated by a high lift coefficient.

As we will discuss in subsequent sections, the lift coefficient for a given


aerodynamic shape is an intrinsic value of the shape itself, the inclination of the
body to the free-stream direction (the angle of attack), the Mach number, and the
Reynolds number.
On the other hand, the data in Fig. 2.6a show an important Reynolds number effect
on Clmax, with higher values of Clmax corresponding to higher Reynolds numbers.
This should be no surprise. The Reynolds number is a similarity parameter in
aerodynamics which governs the nature of viscous flow. The development of
separated flow over the airfoil at high Cl is a viscous flow effect-the viscous
boundary layer literally separates from the surface. Hence we would expect the
value of Clmax to be sensitive to Re; such a sensitivity is clearly seen in Fig. 2.6a.

Un Reynolds más alto significa que la influencia de las fuerzas viscosas que
provocan el desprendimiento de la capa límite (boundary layer) es menor con
relación a las fuerzas inerciales y por eso necesitamos mayores ángulos de
ataque para provocar el desprendimiento.
The drag curve in Figs. 2.6b and 2.9 shows a very flat minimum-the drag
coefficient is at or near its minimum value for a range of angle of attack varying
from -2° to +2°. For this angle-of-attack range, the drag is due to friction drag and
pressure drag. In contrast, the rapid increase in cd which occurs at higher values of
a is due to the increasing region of separated flow over the airfoil, which creates a
large pressure drag.

The variation of Cd with Reynolds number is also shown in Fig. 2,6b. Basic viscous
flow theory and experiments show that the local skin-friction coefficient Cd on a
surface, say, for a flat plate, varies as Cd=1/√Re for laminar flow, and
approximately Cd=1 / (Re )0.2 for turbulent flow. Hence, it is no surprise that Cdmin
in Fig, 2.6b is sensitive to Reynolds number and is larger at the lower Reynolds
numbers. Moreover, the Reynolds number influences the extent and characteristics
of the separated flow region, and hence it is no surprise that Cd at the larger
values of a is also sensitive to the Reynolds number
Also shown in Fig. 2.6b is the variation of the moment coefficient about the
aerodynamic center. By definition, the aerodynamic center is that point on the
airfoil about which the moment is independent of the angle of attack. We discuss
the concept of the aerodynamic center in greater detail in Section 2.6. However,
note that, true to its definition, the experimentally measured value of CMmac in Fig.
2.6b is essentially constant over the range-of-lift coefficient (hence constant over
the range of angle of attack).

We have already defined the aerodynamic center as that point on a body about
which the moments are independent of the angle of attack; that is. practical range
of angle of attack. At first thought, such a seems strange. How can such a point
exist, and how can it be found? We address these questions in this section
The subject of drag has been made confusing historically because so many
different types of drag have been defined and discussed over the years. However,
we can easily cut through this confusion by recalling the discussion in Section 2.2.
There are only two sources of aerodynamic force on a body moving through a fluid-
the pressure distribution and the shear stress distribution acting over the body
surface.
Therefore, there are only two general types of drag:
Pressure drag Due to a net imbalance of surface pressure acting in the drag
direction
Friction drag Due to the net effect of shear stress acting in the drag direction
All the different types of drag that have been defined in the literature fall in one or
the other of the above two categories. It is important to remember this.

Except for the isolated cases of drag due to lift at small angle of attack and
supersonic wave drag for smooth, slender bodies, drag prediction is beyond the
capability of current numerical aerodynamic models.

Skin-friction drag is self-explanatory; it is due to the frictional shear stress acting on


the surface of the airfoil. Pressure drag due to flow separation is caused by the
imbalance of the pressure distribution in the drag direction when the boundary
layer separates from the airfoil surface. (Note that, for an inviscid flow with no flow
separation, theoretically the pressure distribution on the back portion of the airfoil
creates a force pushing forward, which is exactly balanced by the pressure
distribution on the front portion of the airfoil pushing backward. Hence, in a
subsonic inviscid flow over a two-dimensional body, there is no net pressure drag
on the airfoil-this phenomenon is called d'Alembert's paradox after the eighteenth-
century mathematician who first obtained the result In contrast, when the flow
separates from the airfoil, the integrated pressure distribution becomes unbalanced
between the front and back parts of the airfoil, producing a net drag force. This is
the pressure drag due to flow separation.)

Drag: Resistencia o arrastre:


Debido a las fuerzas tangenciales o de corte (shear stress) fricción (cf)
Debido a la distribución de presión y separación del flujo atrás del perfil (form drag
o pressure drag coefficient due flow separation) (cf.p)

For relatively thin airfoils and wings, cf (cf es el coeficiente de fricción) an be


approximated by formulas for a flat plate. But even here there are major
uncertainties in regard to the transition of laminar flow to turbulent flow in the
boundary layer: Turbulence is still a major unsolved problem in classical physics,
and the prediction of where on a surface transition occurs. is uncertain. For a
purely laminar flow, cf for a flat plate in incompressible flow is given by cf=Df/qS

cf =1.328/√Re laminar

Para flujo turbulento es más difícil

cf= 0.42/ln2(0.056 Re) turbulento


However, there remains the question as to where to apply the above formulas,
which is a matter of where transition occurs. Equation (2.27) is valid as long as the
flow is completely laminar. Equations (2.28) and (2.29) are applicable as long as
the flow is completely turbulent.

The latter is a reasonable assumption for most conventional airplanes in subsonic


flight; the flow starts out laminar at the leading edge, but at the high Reynolds
numbers normally encountered in flight, the extent of laminar flow is very small,
and transition usually occurs very near the leading edge-so close that we can
frequently assume that the surface is completely covered with a turbulent boundary
layer. The location at which transition actually occurs on the surface is a function of
a number of variables; suffice it to say that the transition Reynolds number is
350.000 a 1.000.000

To return to Eq. (2.26), the analytical prediction of cd.p, the form drag coefficient, is
still a current research question. No simple equations exist for the estimation of
cd.p, nor does computational fluid dynamics always give the right answer. Instead,
cc1.p is usually found from experiment. [What really happens is that the net profile
drag coefficient cd,p in Eq. (2.26) is measured, such as given in Fig. 2.18, and then
cd,p can be backed out of Eq. (2.26) if a reasonable estimate of cI exists.] At
subsonic speeds below the drag-divergence Mach number, the variation of CJ with
Mach number is very small; indeed, for a first approximation it is reasonable to
assume that cc1 is relatively constant across the subsonic Mach number range.
This is reflected in the left-hand side of Fig. 2.11.
El Factor de eficiencia de Oswald “e”, similar a la 'eficiencia de envergadura', es un
factor de corrección que representa el cambio en la resistencia (inducida) con la
sustentación de un ala tridimensional o un avión, en comparación con un ala ideal que
tiene el mismo alargamiento y una distribución de sustentación elíptica.

La eficiencia Oswald se define para los casos en que el coeficiente de arrastre del ala o
avión tiene una constante de dependencia cuadrática en el coeficiente de sustentación.

Cdt = Cdo + Cdi

Cd es el promedio del coeficiente de arrastre,


Cd0 es el coeficiente de arrastre de cero levantamiento,
Cl es el coeficiente de levantamiento,
e es el número de eficiencia de Oswald
AR es la relación de aspecto de la aeronave

Si e es menor que 1 entonces Cdi comienza a ser muy grande.


Esa es la mejor parte de δ ya que esta en 0.04 o menos. Eso hace que e sea 0.95
o 0.96 casi 1 con lo cual el valor de Cdi será el menor posible. (en elíptica el valor
de δ es cero, el de e es 1 y por tanto Cdi=Cl2/ΠAr

If our objective is to reduce the induced drag, Eq. (2.30) shows us how to do it
First, we want e to be as close to unity as possible. The value of e is always less
than l except for a wing that has a spanwise lift distribution that varies elliptically
over the span, for which e=1. However, as seen in Fig. 2.39, 8 is usually on the
order of 0.05 or smaller for most wings, which means that e varies between 0.95
and 1.0-a relatively minor effect.
Rather, from Eq. (2.30), we see that the aspect ratio plays a strong role; if we can
double the aspect ratio, then we can reduce the induced drag by a factor of 2. The
fact that increasing the aspect ratio reduces the induced drag also makes physical
sense.
Since AR = b2/S for a wing or fixed area, increasing the aspect ratio moves the
wing tips farther from the center of the wing. Since the strength of the induced flow
due to the wing-tip vortices decays with lateral distance from each
vortex, the farther removed the vortices, the weaker the overall induced flow effects
and hence the smaller the indμced drag. Thus, the clear message from Eq. (2.30)
is that increasing the aspect ratio is the major factor in reducing the induced drag.

Eq. (2.30), we see that it makes physical sense that Cd, should be a function of the
lift coefficient (and a strong function, at that, varying as the square of CL). This is
because the generation of wing-tip vortices is associated with a higher pressure
over the bottom of the wing and a lower pressure over the top of the wing-the same
mechanism that produces lift. Indeed, it would be naive for us to assume that lift is
free. The induced drag is the penalty that is paid for the production of lift.
Some special-purpose aircraft have larger aspect ratios. Sailplanes have aspect
ratios that range from 10 to about 30. For example, the Schweizer SGS 1-35 has
an aspect ratio of 23.3. The Lockheed U-2 reconnaissance aircraft (Fig. 2.40) has
as aspect ratio of 14.3 and is capable of flying as high as 90,000 ft. [Reducing the
induced drag for the U-2 was of paramount importance. At very high altitudes,
where the air density is low, the U~2 generates its lift by flying at high values of Cl.
From Eq. (2.30), the induced drag is going to be large. To minimize this effect, the
designers of the U-2 exerted every effort to make the aspect ratio as large as
possible)
Como la velocidad no se puede mejorar y la superficie del ala es chica, la baja de
densidad requiere necesariamente un elevado Cl pára sostener el W del avión U2

L=W= ½ ρ V2 Cl.S

Aspect ratio is one of the most important design features of an airplane. For
subsonic airplane design, it is a major factor in determining the maximum value of
L/D at cruise conditions, which in turn has a major impact on the maximum range
of an airplane ( discussed in Chapter 5). Everything else being equal, the higher
the aspect ratio, the higher the maximum L/D. Of course, in any airplane design
process, not everything else is equal. As noted earlier, as the design aspect ratio is
increased, the wing structure must be made stronger. This increases the weight of
the airplane, which is an undesirable feature. So the airplane designer is faced with
a compromise--one of many in the airplane design process (as discussed in
Chapters 7 and 8). However, the point made here is that, during the interactive
design process, if it becomes important to increase the design value of the
maximum L/D, then one of the powerful tools available to the designer is an
increase in aspect ratio.

Potrebbero piacerti anche