Aerodynamics
Aerodynamics
CHAPTER 2
Applied Aerodynamics
Panel method and Pressure distribution
The panel method is an analysis method that can be used to arrive at an
approximate solution for the forces acting on an object in a flow. The method,
as we present it here, is based on inviscid flow analysis, so it is limited to the
resultant pressure forces over the surface. The panel method is basically a
numerical approximation that relies on using discrete elements on the surface of
an object and then prescribing a flow element (such as a vortex or doublet or
source or sink) on each element that will satisfy certain boundary conditions
(like no flow crosses the surface of the object). The interaction of the elements
are accounted for and must also satisfy the condition that far from the object the
flow should be equal to the free stream velocity approaching the object. There
are a number of books and papers written that describe the method in very
general terms and even the inclusion of viscous forces to some degree. But here
we are just introducing the method to get a feel for its usefulness in external
flows, so we will use a simply geometry with a simply distribution of flow
elements. More complicated models exist but they all are based on the
simplified form presented here.
We will assume that we have potential flow such that the governing equation for
the flow field is the Laplace of the velocity potential, ∇2ϕ=0. The boundary
condition at an impermeable surface, where the velocity normal to the surface is
zero, is ∇ϕ⋅n=0. Also, we can put our frame of reference on the object so fluid
flow approaches the object. Keep in mind that since it is inviscid there may be a
nonzero velocity component tangent to the surface. Also, the goal is to determine
the velocity on the surface, and once this is found the Bernoulli equation can be
used to find the local pressure distribution. The pressure can then be integrated
over the surface to find the force by the fluid flow.
Without deriving this it can be shown that the following defines the velocity
potential at any point P in the flow field (using Green’s Identify):
where the integral is over the surface area of the flow (assuming two
dimensional flow), S. This equation indicates that to solve for the velocity
potential we must evaluate the integral on the flow boundaries (both the solid
surface and infinitely far away).
All this is nice and can be a powerful tool to find ϕ and therefore the velocity in
the flow. But we really don’t need to find the potential of the entire flow, what our
goal is, is to set up a flow that we know satisfies the velocity boundary condition at
the surface of the object (with no velocity component across the surface) and far
from the surface where the velocity is known to be the freestream velocity. Once
this is established the force can be found. That is to say we only need to evaluate
the surface velocity and then the pressure on the surface.
The general approach is to select a “grid” which is a series of “panels” that form
the surface. Here we take the panels as straight flat surfaces arranged over the
real surface. In the limit of very small panels the constructed surface will
simulate the actual surface. On each panel we place a distribution of flow
elements (like sources, sinks, vortices, etc.) that when combined together will
result in a flow field that will satisfy the surface boundary condition. There are
lots of ways to identify which elements to use and how they may be distributed
on the panels. Here we will use vortex elements, with one placed on each panel.
The net flow is the result of superposition of the flow set up by each vortex on
each panel element. So at each point in the field we add together the flow
caused by all of the panel elements using the superposition rule. The panels that
are far away from a given point will have less and less influence on the flow
because the strength of the flow caused by a flow element decreases with
distance from the element origin. For instance for a “source” the velocity
decreases with increasing radial position because the flow is spreading out away
from the source. But the influence never really goes to zero.
In placing a series of panels over the surface we first need to specify the size of
each panel. We place a vortex of some strength somewhere on the panel (whose
location is shown below) and we must identify points on the surface where we
want to make sure that the velocity is zero across the surface. To be clear,
individual points on each panel are used to evaluate the element (vortex) flow
field –it needs its own origin, or coordinate system, to write an equation for the
flow generated by this element. We also only pick a point on the panel to check
to make sure that the net sum of contributions from all elements results in zero
flow across the surface. The fact that we only satisfy the condition at one point
on each panel will be satisfactory if the panels are made to be reasonably small.
These points are called “collocation points” on each panel. In the end each panel
will have coordinates that define its location on the surface, coordinates for the
element location on the surface and coordinates for the collocation point.
he method of solution for the force on the object is the determination of the
magnitudes or strengths of the elements on each panel. Once we have this
distribution of strengths we can calculate the total lift on the surface that
results from all of these elements. Recall that the lift experienced by a
surface is really the component of the pressure at the surface integrated
over the surface area in a direction normal to the approach velocity vector
of the flow — it is not necessarily vertical, but normal to the freestream
velocity.
For illustration we are going to use flat plate panels with vortex elements,
one per panel. To get an idea of how to define the vortex origins on each
panel we can examine flow over a single flat plate at some angle of attack
to the freestream velocity vector, α.
Stream function
We see that the equation for a streamline is given by setting the stream
function equal to a constant (i.e., c1, c2, c3, etc.). Two different streamlines
are illustrated in Figure ; streamlines ab and cd are given by ψ = c1 and
ψ = c2, respectively using arbitrary constant of integration c. Let us define
the stream function more precisely in order to reduce this arbitrariness. Let
us define the numerical value of ψ such that the difference ψ between ψ =
c2 for streamline cd and ψ = c1 for streamline ab is equal to the mass flow
between the two streamlines. Since Figure is a two-dimensional flow, the
mass flow between two streamlines is defined per unit depth perpendicular
to the page. That is, in Figure we are considering the mass flow inside a
stream tube bounded by streamlines ab and cd, with a rectangular cross-
sectional area equal to _n times a unit depth perpendicular to the page.
Here, _n is the normal distance between ab and cd, as shown in Figure .
Hence, mass flow between streamlines ab and cd per unit depth
perpendicular to the page is
Kutta condition
The Kutta condition is a principle in steady-flow fluid dynamics, especially
aerodynamics, that is applicable to solid bodies with sharp corners, such as the
trailing edges of airfoils. A body with a sharp trailing edge which is moving
through a fluid will create about itself a circulation of sufficient strength to hold
the rear stagnation point at the trailing edge. In fluid flow around a body with a
sharp corner, the Kutta condition refers to the flow pattern in which fluid
approaches the corner from both directions, meets at the corner, and then flows
away from the body. None of the fluid flows around the corner, remaining
attached to the body. When a smooth symmetric body, such as a cylinder with
oval cross-section, moves with zero angle of attack through a fluid it generates
no lift. There are two stagnation points on the body - one at the front and the
other at the back. If the oval cylinder moves with a non-zero angle of attack
through the fluid there are still two stagnation points on the body - one on the
underside of the cylinder, near the front edge; and the other on the topside of the
cylinder, near the back edge. The circulation around this smooth cylinder is zero
and no lift is generated, despite the positive angle of attack.
The Kutta condition allows an aerodynamicist to incorporate a significant effect
of viscosity while neglecting viscous effects in the underlying conservation of
momentum equation. It is important in the practical calculation of lift on a wing.
The equations of conservation of mass and conservation of momentum applied
to an inviscid fluid flow, such as a potential flow, around a solid body result in
an infinite number of valid solutions. One way to choose the correct solution
would be to apply the viscous equations, in the form of the Navier–Stokes
equations. However, these normally do not result in a closed-form solution. The
Kutta condition is an alternative method of incorporating some aspects of
viscous effects, while neglecting others, such as skin friction and some other
boundary layer effects. The condition can be expressed in a number of ways.
One is that there cannot be an infinite change in velocity at the trailing edge.
Although an inviscid fluid can have abrupt changes in velocity, in reality
viscositysmooths out sharp velocity changes. If the trailing edge has a non-zero
angle, the flow velocity there must be zero. At a cusped trailing edge, however,
the velocity can be non-zero although it must still be identical above and below
the airfoil. Another formulation is that the pressure must be continuous at the
trailing edge. The Kutta condition does not apply to unsteady flow.
Experimental observations show that the stagnation point (one of two points on
the surface of an airfoil where the flow speed is zero) begins on the top surface
of an airfoil (assuming positive effective angle of attack) as flow accelerates
from zero, and moves backwards as the flow
accelerates. Once the initial transient effects have died out, the stagnation point
is at the trailing edge as required by the Kutta condition.
The velocity induced vertically (vi) at any point on the mean line can be found
by summing up the effects of small individual segments (ds) of the vorticity
distribution. where (x) is the location at which the induced velocity is being
calculated and (s) is the chord-wise location of the vortex element. On
integration after substitution for s,x and γ this gives,
On a three-dimensional, finite wing, lift over each wing segment (local lift per
unit span, or does not correspond simply to what two-dimensional analysis
predicts. Instead, this local amount of lift is strongly affected by the lift
generated at neighboring wing sections. As such, it is difficult to predict
analytically the overall amount of lift that a wing of given geometry will
generate. The lifting-line theory yields the lift distribution along the span-wise
direction, based only on the wing geometry (span-wise distribution of chord,
airfoil, and twist) and flow conditions.
Aerodynamic potential flow codes or panel codes are used to determine the
fluid velocity, and subsequently the pressure distribution, on an object. This
Irrotational ∇ × V = 0
However, the incompressible flow assumption may be removed from the potential
flow derivation leaving:
Potential Flow (inviscid, irrotational, steady)
oscillating body is an example of a Stokes boundary layer, while the Blasius boundary layer
refers to the well-known similarity solution near an attached flat plate held in an oncoming
unidirectional flow and Falkner–Skan boundary layer, a generalization of Blasius profile.
When a fluid rotates and viscous forces are balanced by the Coriolis effect (rather than
convective inertia), an Ekman layer forms. In the theory of heat transfer, a thermal boundary
layer occurs. A surface can have multiple types of boundary layer simultaneously. The
viscous nature of airflow reduces the local velocities on a surface and is responsible for skin
friction. The layer of air over the wing's surface that is slowed down or stopped by viscosity
is the boundary layer. There are two different types of boundary layer flow: laminar and
turbulent.
Laminar boundary layer
The laminar boundary is a very smooth flow, while the turbulent boundary layer contains
swirls or "eddies." The laminar flow creates less skin friction drag than the turbulent flow,
but is less stable. Boundary layer flow over a wing surface begins as a smooth laminar flow.
Transition
The process of a laminar flow becoming turbulent is known as laminar-turbulent
transition. This is an extraordinarily complicated process which at present is not fully
understood. However, as the result of many decades of intensive research, certain
features have become gradually clear, and it is known that the process proceeds
through a series of stages. "Transitional flow" can refer to transition in either
direction that is laminar-turbulent transitional or turbulent-laminar transitional flow.
Prof. Zoya Rizvi (NMIMS University)
2022
Course: Aerodynamics in Flight Design
Consider a stationary body with a fluid flowing around it, like the semi-infinite flat
plate with air flowing over the top of the plate (assume the flow and the plate
extends to infinity in the positive/negative direction perpendicular to the (x-y plane).
At the solid walls of the body the fluid satisfies a no-slip boundary condition and has
zero velocity, but as you move away from the wall, the velocity of the flow
asymptotically approaches the free stream mean velocity. Therefore, it is impossible
to define a sharp point at which the boundary layer becomes the free stream, yet
this layer has a well-defined characteristic thickness. The parameters below provide a
useful definition of this characteristic, measurable thickness. Also included in this
boundary layer description are some parameters useful in describing the shape of
the boundary layer.
This dictates the inviscid pressure distribution on the cylinder which is shown by a
firm line in Fig.
Log Law:
In fluid dynamics, the law of the wall (also known as the logarithmic law of the wall) states
that the average velocity of a turbulent flow at a certain point is proportional to the
logarithm of the distance from that point to the "wall", or the boundary of the fluid
region.