Download as pdf or txt
Download as pdf or txt
You are on page 1of 81

Fluids – Lecture 1 Notes

1. Introductory Concepts and Definitions


2. Properties of Fluids

Reading: Anderson 1.1 (optional), 1.2, 1.3, 1.4

Introductory Concepts and Definitions


Fluid Mechanics and Fluid Dynamics encompass a huge range of topics which deal with
the behavior of gasses and liquids. In UE we will focus mainly on the topic subset called
Aerodynamics, with a bit af Aerostatics in the beginning.
Merriam Webster’s definitions:

Aerostatics: a branch of statics that deals with the equilibrium of gaseous fluids and
of solid bodies immersed in them
Aerodynamics: a branch of dynamics that deals with the motion of air and other
gaseous fluids and with the forces acting on bodies in motion relative to such fluids

Related older terms are Hydrostatics and Hydrodynamics, usually used for situtations in-
volving liquids. There is surprisingly little fundamental difference between the Aero– and
Hydro– disciplines. They differ mainly in the applications (e.g. airplanes vs. ships).

Difference between a Solid and a Fluid (Liquid or Gas):

Solid: Applied tengential force/area (or shear stress) τ produces a proportional de-
formation angle (or strain) θ.
τ = Gθ
The constant of proportionality G is called the elastic modulus, and has the units of
force/area.

Fluid: Applied shear stress τ produces a proportional continuously-increasing defor-


mation (or strain rate) θ̇.
τ = µθ̇
The constant of proportionality µ is called the viscosity, and has the units of
force×time/area.

τ stress τ stress

.
θ strain θ strain
rate

Solid Fluid

Properties of Fluids
Continuum vs molecular description of fluid
Liquids and gases are made up of molecules. Is this discrete nature of the fluid important
for us? In a liquid, the answer is clearly NO. The molecules are in contact as they slide past
each other, and overall act like a uniform fluid material at macroscopic scales.
In a gas, the molecules are not in immediate contact. So we must look at the mean free
path, which is the distance the average molecule travels before colliding with another. Some
known data for air:

Mean free path at 0 km (sea level) : 0.0001 mm

Mean free path at 20 km (U2 flight) : 0.001 mm

Mean free path at 50 km (balloons) : 0.1 mm

Mean free path at 150 km (low orbit) : 1000 mm = 1m

The mean free path is vastly smaller than the typical dimension of any atmospheric vehicle.
So even though the lift on a wing is due to the impingement of discrete molecules, we
can assume the air is a continuum for the purpose of computing this lift. In contrast,
computing the slight air drag on an orbiting satellite requires treating the air as discrete
isolated particles.

Pressure
Pressure p is defined as the force/area acting normal to a surface. A solid surface doesn’t
actually have to be present. The pressure can be defined at any point x, y, z in the fluid, if
we assume that a infinitesimally small surface ΔA could be placed there at whim, giving a
resulting normal force ΔFn .
ΔFn
p = lim
ΔA→0 ΔA
The pressure can vary in space and possibly also time, so the pressure p(x, y, z, t) in general
is a time-varying scalar field.

ΔFn

ΔΑ

Normal force on area element due to pressure

Density
Density ρ is defined as the mass/volume, for an infinitesimally small volume.
Δm
ρ = lim
ΔV→0 ΔV

Like the pressure, this is a point quantity, and can also change in time. So ρ(x, y, z, t) is also
a scalar field.

2
Velocity
We are interested in motion of fluids, so velocity is obviously important. Two ways to look
at this:

• Body is moving in stationary fluid – e.g. airplane in flight

• Fluid is moving past a stationary body – e.g. airplane in wind tunnel

The pressure fields and aerodynamics forces in these two cases will be the same if all else is
equal. The governing equations we will develop are unchanged by a Galilean Transformation,
such as the switch from a fixed to a moving frame of reference.
Consider a fluid element (or tiny “blob” of fluid) as it moves along. As it passes some point
B, its instantaneous velocity is defined as the velocity at point .B
~ at a point = velocity of fluid element as it passes that point
V

This velocity is a vector, with three separate components, and will in general vary between
different points and different times.
~ (x, y, z, t) = u(x, y, z, t) ı̂ + v(x, y, z, t) ˆ + w(x, y, z, t) kˆ
V
~ is a time-varying vector field, whose components are three separate time-varying scalar
So V
fields u, v, w.
B V
fluid
element

e
streamlin

Velocity at point B

A useful quantity to define is the speed , which is the magnitude of the velocity vector.
� � √
~ �� = u2 + v 2 + w 2
V (x, y, z, t) = ��V

In general this is a time-varying scalar field.

Steady and Unsteady Flows


If the flow is steady, then p, ρ, V~ don’t change in time for any point, and hence can be given
~ (x, y, z). If the flow is unsteady, then these quantities do change in
as p(x, y, z), ρ(x, y, z), V
time at some or all points.
For a steady flow, we can define a streamline, which is the path followed by some chosen
fluid element. The figure above shows three particular streamlines.

3
Fluids – Lecture 2 Notes
1. Hydrostatic Equation
2. Manometer
3. Buoyancy Force

Reading: Anderson 1.9

Hydrostatic Equation
Consider a fluid element in a pressure gradient in the vertical y direction. Gravity is also
present.
y

dp dx
dy dy

dz
x

g
z

If the fluid element is at rest, the net force on it must be zero. For the vertical y-force in
particular, we have

Pressure force + Gravity force = 0


� �
dp
p dA − p + dy dA − ρ g dV = 0
dy
dp
− dy dA − ρ g dV = 0
dy
The area on which the pressures act is dA = dx dz, and the volume is dV = dx dy dz, so that
dp
− dx dy dz − ρ g dx dy dz = 0
dy
dp = −ρg dy (1)

which is the differential form of the Hydrostatic Equation. If we make the further assumption
that the density is constant, this equation can be integrated to the equivalent integral form.

p(y) = p0 − ρgy (2)

The constant of integration p0 is the pressure at the particular location y = 0. Note that
this integral form is valid provided the density is constant within the region of interest
.

Application to a Manometer
A manometer is a U-shaped tube partially filled with a liquid, as shown in the figure. Two
different pressures p1 and p2 are applied to the two legs of the tube, causing the two liquid
columns to have different heights h1 and h2 .
y
p1 p2

h1

h2

p0

We now pick p0 to be the pressure at some point of the tube (at the bottom for instance),
and apply equation (2) to each leg of the tube.
p1 = p0 − ρgh1
p2 = p0 − ρgh2
Subtracting these two equations then gives the difference of the pressures in terms of the
liquid height difference.
p2 − p1 = ρg(h1 − h2 ) (3)
If tube 1 is left open to the atmosphere, so that p1 = patm , then p2 can be measured simply by
applying it to tube 2, measuring the height difference Δh = h1 − h2 , and applying equation
(3) above.

p2 = patm + ρg Δh

This requires knowing the density ρ of the fluid to sufficient accuracy.

Buoyancy
Now consider an object of arbitrary shape immersed in the pressure gradient. The ob-
ject’s volume can be divided into vertical “matchstick” volumes, each of infinitesimal cross-
sectional area dA = dx dz, and finite height Δh.

The vertical y-direction pressure force on each volume is


� �
dp
dF = p dA − p + Δh dA
dy
dp
dF = − Δh dA
dy
dF = ρg dV

2
y
dA = dx dz

dp Δh
dy

g
z
where dp/dy has been replaced by −ρg using the Hydrostatic Equation (1), and the volume
of the infinitesimal volume is Δh dA = dV. Integrating the last equation above then gives
the total buoyancy force on the object.

F = ρgV

It is important to note that V is the overall volume of the object, while ρ is the density of
the fluid. The product ρV is recognized as the mass of the fluid displaced by the object, and
ρgV is the corresponding weight, giving the well known Archimedes Principle:

Buoyancy force on body = Weight of fluid displaced by body

Fluids – Lecture 3 Notes


1. 2-D Aerodynamic Forces and Moments
2. Center of Pressure
3. Nondimensional Coefficients
Reading: Anderson 1.5 – 1.6

Aerodynamics Forces and Moments


Surface force distribution
The fluid flowing about a body exerts a local force/area (or stress) f~ on each point of the
body. Its normal and tangential components are the pressure p and the shear stress τ .

f f
τ p

r r ds

local pressure and shear stress components


( τ magnitude greatly exaggerated)
force/area distribution on airfoil
N
L R
α
R
M
α
D
V
A
resultant force, and moment about ref. point alternative components of resultant force

The figure above greatly exaggerates the magnitude of the τ stress component just to make
it visible. In typical aerodynamic situations, the pressure p (or even the relative pressure
p − p∞ ) is typically greater than τ by at least two orders of magnitude, and so f~ is very
nearly perpendicular to the surface. But the small τ often significantly contributes to drag,
so it cannot be neglected entirely.
The stress distribution f~ integrated over the surface produces a resultant force R,
~ and also
a moment M about some chosen moment-reference point. In 2-D cases, the sign convention
for M is positive nose up, as shown in the figure.

Force components
~ has perpendicular components along any chosen axes. These axes are
The resultant force R
arbitrary, but two particular choices are most useful in practice.

1
Freestream Axes: The R ~ components are the drag D and the lift L, parallel and perpendic-
~∞ .
ular to V
Body Axes: The R ~ components are the axial force A and normal force N, parallel and
perpendicular to the airfoil chord line.
If one set of components is computed, the other set can then be obtained by a simple axis
transformation using the angle of attack α. Specifically, L and D are obtained from N and
A as follows.

L = N cos α − A sin α
D = N sin α + A cos α

Force and moment calculation


A cylindrical wing section of chord c and span b has force components A and N, and mo-
ment M. In 2-D it’s more convenient to work with the unit-span quantities, with the span
dimension divided out.

A′ ≡ A/b N ′ ≡ N/b M ′ ≡ M/b

θ pu c
y
pu (su )
τu
dsu b
su θ τu (su)
dsu
dsl x
α sl
V
τl (sl )
pl (sl )
θ
On the upper surface, the unit-span force components acting on an elemental area of width
dsu are

dNu′ = (−pu cos θ − τu sin θ) dsu


dA′u = (−pu sin θ + τu cos θ) dsu

And on the lower surface they are

dNℓ′ = (pℓ cos θ − τℓ sin θ) dsℓ


dA′ℓ = (pℓ sin θ + τℓ cos θ) dsℓ

Integration from the leading edge to the trailing edge points produces the total unit-span
forces.
TE TE
� �
N′ = dNu′ + dNℓ′
�LE
TE
� LE
TE

A = dAu′ + dAℓ′
LE LE

The moment about the origin (leading edge in this case) is the integral of these forces,
weighted by their moment arms x and y, with appropriate signs.
TE TE TE TE
� � � �

MLE = −x dNu′ + −x dNℓ′ + y dAu′ + y dAℓ′
LE LE LE LE

From the geometry, we have


dy
ds cos θ = dx ds sin θ = −dy = − dx
dx
which allows all the above integrals to be performed in x, using the upper and lower shapes
of the airfoil yu (x) and yℓ (x). Anderson 1.5 has the complete expressions.
Simplifications
In practice, the shear stress τ has negligible contributions to the lift and moment, giving the
following simplified forms.
� �
c dyℓ dyu c
� �

L = cos α (pℓ − pu ) dx + sin α pℓ − pu dx
0 0 dx dx
� c� � � � ��
′ dyu dyℓ
MLE = pu x + y u − pℓ x + yℓ dx
0 dx dx

A somewhat less accurate but still common simplification is to neglect the sin α term in L′ ,
and the dy/dx terms in M ′ .
� c

L ≃ (pℓ − pu ) dx
�0c

MLE ≃ −(pℓ − pu ) x dx
0

The shear stress τ cannot be neglected when computing the drag D ′ on streamline bodies
such as airfoils. This is because for such bodies the integrated contributions of p toward D ′
tend to mostly cancel, leaving the small contribution of τ quite significant.

Center of Pressure
Definition
The value of the moment M ′ depends on the choice of reference point. Using the simplified
form of the MLE integral, the moment Mref for an arbitrary reference point xref is
� c

Mref = −(pℓ − pu ) (x − xref ) dx
0

= MLE + L′ xref

This can be positive, zero, or negative, depending on where xref is chosen, as illustrated in
the figure.
At one particular reference location xcp , called the center of pressure
, the moment is defined
to be zero.
′ ′
Mcp = MLE + L′ xcp ≡ 0

xcp = −MLE /L′

pl −pu

L L L
xcp
or or

M<0 M=0 M<0

The center of pressure asymptotes to +∞ or −∞ as the lift tends to zero. This awkward
situation can easily occur in practice, so the center of pressure is rarely used in aerodynamics
work.
For reasons which will become apparent when airfoil theory is studied, it is advantageous
to define the “standard” location for the moment reference point of an airfoil to be at its
quarter-chord location, or xref = c/4. The corresponding standard moment is usually written
without any subscripts.
� c
′ ′
Mc/4 ≡ M = −(pℓ − pu ) (x − c/4) dx
0

Aerodynamic Conventions
As implied above, the aerodynamicist has the option of picking any reference point for the
moment. The lift and the moment then represent the integrated pl −pu distribution. Consider
two possible representations:

1. A resultant lift L′ acts at the center of pressure x = xcp . The moment about this
point is zero by definition: Mcp

= 0. The xcp location moves with angle of attack in a
complicated manner.

2. A resultant lift L′ acts at the fixed quarter-chord point x = c/4. The moment about
this point is in general nonzero: Mc/4

6= 0.

The figure shows how the L′ , M ′ , and xcp change with angle of attack for a typical cambered
airfoil. Note that with representation 1, the xcp location moves off the airfoil and tends to
+∞ as L′ approaches zero. Fixing the moment reference point, as in representation 2, is a
simpler and preferable approach. Choosing the quarter-chord location for this is especially
attractive, since M ′ then shows little or no dependence on the angle of attack. This surprising
fact will come from a more detailed airfoil analysis later in the course.

4
pl −pu

α=5
α=0
α = −5

1. L’

(M’=0)

xcp xcp xcp

2. L’

M’

c/4 c/4 c/4

Nondimensional Coefficients
The forces and moment depend on a large number of geometric and flow parameters. It
is often advantageous to work with nondimensionalized forces and moment, for which most
of these parameter dependencies are scaled out. For this purpose we define the following
reference parameters:
Reference area: S
Reference length: ℓ
Dynamic pressure: q∞ = 21 ρV∞2
The choices for S and ℓ are arbitrary, and depend on the type of body involved. For aircraft,
traditional choices are the wing area for S, and the wing chord or wing span for ℓ. The
nondimensional force and moment coefficients are then defined as follows:
L
Lift coefficient: CL ≡
q∞ S
D
Drag coefficient: CD ≡
q∞ S
M
Moment coefficient: CM ≡
q∞ Sℓ
For 2-D bodies such as airfoils, the appropriate reference area/span is simply the chord c, and
the reference length is the chord as well. The local coefficients are then defined as follows.
L′
Local Lift coefficient: cℓ ≡
q∞ c
D′
Local Drag coefficient: cd ≡
q∞ c
M′
Local Moment coefficient: cm ≡
q∞ c2
These local coefficients are defined for each spanwise location on a wing, and may vary across
the span. In contrast, the CL , CD , CM are single numbers which apply to the whole wing.

Fluids – Lecture 4 Notes


1. Dimensional Analysis – Buckingham Pi Theorem
2. Dynamic Similarity – Mach and Reynolds Numbers
Reading: Anderson 1.7

Dimensional Analysis
Physical parameters
A large number of physical parameters determine aerodynamic forces and moments. Specif-
ically, the following parameters are involved in the production of lift.
Parameter Symbol Units
Lift per span L′ mt−2

Angle of attack α —

Freestream velocity V∞ lt−1

Freestream density ρ∞ ml−3

Freestream viscosity µ∞ ml−1 t−1

Freestream speed of sound a∞ lt−1

Size of body (e.g. chord) c l

For an airfoil of a given shape, the lift per span in general will be a function of the remaining
parameters in the above list.

L′ = f (α, ρ∞ , V∞ , c, µ∞ , a∞ ) (1)

In this particular example, the functional statement has N = 7 parameters, expressed in a


total of K = 3 units (mass m, length l, and time t).

Dimensionless Forms
The Buckingham Pi Theorem states that this functional statement can be rescaled into an
equivalent dimensionless statement

Π1 = f¯( Π2 , Π3 . . . ΠN −K )

having only N −K = 4 dimensionless parameters. These are called Pi products, since they
are suitable products of the dimensional parameters. In the particular case of statement (1),
suitable Pi products are:
L′
Π1 = 1 = cℓ lift coefficient
ρ V2c
2 ∞ ∞
Π2 = α = α angle of attack
ρ∞ V∞ c
Π3 = = Re Reynolds number
µ∞
V∞
Π4 = = M∞ Mach number
a∞
The dimensionless form of statement (1) then becomes

¯ Re, M∞ )
cℓ = f(α, (2)

1
We see that the original 6 dimensional parameters which influence L′ has been reduced to
only 3 dimensionless parameters which influence cℓ .

Benefits of non-dimensionalization
The reduction of parameter count is potentially a huge simplification. Consider an exaustive
lift-measurement experiment where the effect of all parameters is to be determined. Let’s
assume that in this experiment we need to give each parameter 5 distinct values in order
to adequately ascertain its effect on the lift. If we work with the 6 dimensional parameters
in statement (1), then the number of possible parameter combinations and experimental
runs required is 56 = 15625 (!). But if we work with the 3 dimensionless parameters in
statement (2), the number of parameter combinations and experimental runs is only 53 = 125,
which is more than a hundredfold reduction in effort. Nondimensionalization is clearly a
powerful technique for minimizing experimental effort.
The benefits of non-dimensionalization also extend to theoretical work. Deducing a statement
such as (2) at the outset can be useful to guide subsequent detailed analysis. Theoretical
results are also usually more concise and clear when presented in dimensionless form.

Derivation of dimensionless forms


Anderson 1.7 has details on how the Pi product combinations can be derived for any complex
situation using linear algebra. In many cases, however, the products can be obtained by
physical insight, or perhaps by inspection. Several rules can be applied here:
• Any parameter which is already dimensionless, such as α, is automatically one of the
Pi products.
• If two parameters have the same units, such as V∞ and a∞ , then their ratio (M∞ in
this case) will be one of the Pi products.
• A power or simple multiple of a Pi product is an acceptable alternative Pi product. For
example, (V∞ /a∞ )2 is an acceptable alternative to V∞ /a∞ , and ρ∞ V∞2 is an acceptable
alternative to 12 ρ∞ V∞2 . Which particular forms are used is a matter of convention.
• Combinations of Pi products can replace the originals. For example, the 3rd and 4th
products in the example could have been defined as
ρ∞ a∞ c
Π3 = = Re/M∞
µ∞
V∞
Π4 = = M∞
a∞
which is workable alternative, but perhaps less practical, and certainly less traditional.

Dynamic Similarity
It is quite possible for two differently-sized physical situations, with different dimensional
parameters, to nevertheless reduce to the same dimensionless description. The only require-
ment is that the corresponding Pi products have the same numerical values.

2
Airfoil flow example
Consider two airfoils which have the same shape and angle of attack, but have different sizes
and are operating in two different fluids. Let’s omit the ()∞ subscript for clarity.

Airfoil 1 (sea level) Airfoil 2 (cryogenic tunnel)


α1 = 5◦ α2 = 5◦
V1 = 210m/s V2 = 140m/s
ρ1 = 1.2kg/m3 ρ2 = 3.0kg/m3
µ1 = 1.8 × 10−5 kg/m-s µ2 = 1.5 × 10−5 kg/m-s
a1 = 300m/s a2 = 200m/s
c1 = 1.0m c2 = 0.5m

Airfoil 1 − Sea level air Airfoil 2 − Cryogenic tunnel

The Pi products evaluate to the following values.


Airfoil 1 Airfoil 2
α1 = 5◦
α2 = 5◦
Re1 = 1.4 × 107
Re2 = 1.4 × 107
M1 = 0.7
M2 = 0.7
Since these are also the arguments to the f¯ function, we conclude that the cℓ values will be
the same as well.

f¯(α1 , Re1 , M1 ) = f¯(α2 , Re2 , M2 )


cℓ 1 = cℓ 2

When the nondimensionalized parameters are equal like this, the two situations are said to
have dynamic similarity. One can then conclude that any other dimensionless quantity must
also match between the two situations. This is the basis of wind tunnel testing, where the
flow about a model object duplicates and can be used to predict the flow about the full-
size object. The prediction is correct only if the model and full-size objects have dynamic
similarity.

Fluids – Lecture 5 Notes


1. Effects of Reynolds Number
2. Effects of Mach Number
Reading: Anderson 1.10, 1.11

Dimensional Analysis has identified the important dimensionless parameters (or Π products),
which determine the nature of a given aerodynamic flow situation. The main parameters are
the Reynolds number and the Mach number, and their values in any given flow situation will
determine the nature of the flow (strongly viscous vs. nearly inviscid, low speed vs. high
speed)

Effects of Reynolds Number


Interpretation
The Reynolds number can be interpreted as a comparison of the pressure forces and viscous
shear forces which act on the body, done by forming their ratio.
pressure forces p − p∞ ρu2 ρ∞ V∞2 ρ∞ V∞ c
= ∼ ∼ ∼ ≡ Re∞
shear forces τ µ du/dn µ∞ V∞ /c µ∞
Note that p − p∞ is considered rather than p itself, since p∞ is just a large constant bias
which produces no net force on a body.
The Reynolds number is therefore a measure of the importance of pressure forces relative
to viscous shear forces. Hence, if the Reynolds number increases, the viscous effects on the
flow get progressively less important.
The viscosity of air and water is quite small when expressed in common units.
Air @ STP Water @ 15◦ C
µ 1.78 × 10−5 kg/m-s 1.15 × 10−3 kg/m-s

µ/ρ = ν 1.45 × 10−5 m2 /s 1.15 × 10−6 m2 /s

The ratio µ/ρ = ν is called the kinematic viscosity. This is more relevant than µ itself, since
the ratio is what actually appears in the Reynolds number.
V∞ c
Re∞ =
ν∞

From the small values of ν in the table above, it is clear that typical aerodynamic and
hydrodynamic flows will have very large Reynolds numbers. This can be seen in the following
table, which gives the Reynolds numbers based on the chord length of common winged
objects.

Object Re
Butterfly 5K
Pigeon 50 K
RC glider 100 K
Sailplane 1M
Business jet 10 M
Boeing 777 50 M

The Reynolds number is large even for insects, which means that the flow can be assumed
to be inviscid (i.e. µ = 0 and τ = 0) almost everywhere. The only place where the viscous
shear is significant is in boundary layers which form adjacent to solid surfaces and become a
wake trailing downstream.
boundary layer boundary layer

Re = 10 K Re = 1 M

cd
0.035 cd 0.0045
The larger the Reynolds number is, the thinner the boundary layers are relative to the size
of the body, and the more the flow behaves as though it was inviscid. This has a number of
consequences of great practical significance:

1. Neglecting the viscous shear stress greatly simplifies the equations of fluid motion, al-
lowing them to be manipulated and solved much more easily. Simple solutions then give
better insight and understanding of aerodynamic behavior of bodies such as airfoils,
wings, propellers, turbine blades, etc.
2. If the Reynolds number is large enough (i.e. viscosity is small enough), then its effects
on some aerodynamic forces and moments can be neglected. For example, if Re > 106
or so, then it has almost no influence on the aerodynamic lift and moment, which
means that the general dependencies for cℓ and cm simplify as follows.
cℓ = cℓ (α, M∞ , Re∞ ) cℓ = cℓ (α, M∞ )
→ (Re∞ large)
cm = cm (α, M∞ , Re∞ ) cm = cm (α, M∞ )

Such a simplification is not appropriate for cd (α, M∞ , Re∞ ), however.

Streamlined vs. Bluff Bodies


The above assumption that viscosity has negligible effects on lift and moment is valid only
for streamline bodies which gradually taper to a point or sharp edge, and are roughly
aligned with the freestream flow direction. Examples are slender bodies, airfoils, and wings,
operating at small angles of attack. For bluff bodies which have blunt rear faces, or for
streamline bodies at large angles of attack, viscosity always plays an important role for all
the forces and moments. Such flows exhibit flow separation , which is the sudden thickening
or breakaway of the boundary layer from the surface, resulting in a thick trailing wake.
flow separation
flow separation

Stalled Streamline body Bluff body


cd 0.1 − 0.5 cd 0.6 − 1.2
Bodies with separation exhibit much larger drag forces than comparably-sized streamline

bodies without separation. When an airfoil stalls, its drag will typically increase by a factor

of 10 or more. Most of this drag increase is due to a sharp rise of pressure drag, defined as
the integrated pressure p − p∞ normal to the body. The friction drag, which is the integrated
tangential shear stress τ , is not significantly different between streamline and bluff bodies of
similar size.

Drag reference length


The reference length used to define cd is arbitrary. Traditionally, for streamline bodies such
as an airfoil we use the streamwise length c. For bluff bodies we use the transverse height,
typically denoted by h. For a circular cylinder they are of course the same, and equal to the
diameter d.

Effects of Mach Number


Interpretation
The Mach number is defined as the ratio of the freestream speed to the speed of sound.
V∞
≡ M∞
a∞
The speed of sound a∞ is the velocity at which pressure disturbances travel in the fluid,
although what this implies for the flow is not immediately apparent.
Another interpretation of the Mach number arises when we estimate the fractional variation
of pressure in the flowfield.
1
Δp q∞ ρ V2
2 ∞ ∞ γ ρ∞ V
∞2 γ V
∞2 γ 2
∼ = = = = M∞
p∞ p∞ p∞ 2 γp∞ 2 a∞
2 2

where γ = cp /cv is the thermodynamic ratio of specific heats. So the Mach number squared
is a measure of the fractional pressure variations in the flowfield, as shown in the figure.

p/p

1
Δp / p ~ M2

Low-Speed flows
A low-speed flow is defined as one for which
2
M∞ ≪ 1.

From the above derivation this implies that fractional pressure changes are small, and from
thermodynamics we then conclude that that density and temperature changes must also be
small.
Δp Δρ ΔT
≪1 , ≪1 , ≪1
p ρ T
So an important feature of a low-speed flow is nearly constant fluid properties throughout the
flow field. All hydrodynamic flows can be classified as low-speed flows, since liquids feature
an essentially constant density.
One practical feature of low-speed flows is that they are extremely insensitive to Mach
2
number variations, as long as M∞ ≪ 1. So a flow with M∞ = 0.01 is almost indistinguishible
from a flow with M∞ = 0.1 when compared using dimensionless variables. So for low speed
flows we can remove the influence of M∞ from the dependency list.

cℓ = cℓ (α, M∞ ) cℓ = cℓ (α) � �
2
cm = cm (α, M∞ ) → cm = cm (α) M∞ ≪1
cd = cd (α, M∞ , Re∞ ) cd = cd (α, Re∞ )

This is a considerable simplification.

4
Fluids – Lecture 6 Notes
1. Control Volumes
2. Mass Conservation
3. Control Volume Applications
Reading: Anderson 2.3, 2.4

Control Volumes
In developing the equations of aerodynamics we will invoke the firmly established and time-
tested physical laws:

Conservation of Mass

Conservation of Momentum

Conservation of Energy

Because we are not dealing with isolated point masses, but rather a continuous deformable
medium, we will require new conceptual and mathematical techniques to apply these laws
correctly.
One concept is the control volume, which can be either finite or infinitesimal. Two types of
control volumes can be employed:
1) Volume is fixed in space (Eulerian type). Fluid can freely pass through the volume’s
boundary.
2) Volume is attached to the fluid (Lagrangian type). Volume is freely carried along with the
fluid, and no fluid passes through its boundary. This is essentially the same as the free-body
concept employed in solid mechanics.

V V

Eulerian Control Volume Lagrangian Control Volume


fixed in space moving with fluid

Both approaches are valid. Here we will focus on the fixed control volume.

Mass Conservation
Mass flow
Consider a small patch of the surface of the fixed, permeable control volume. The patch has

area A, and normal unit vector n̂. The plane of fluid particles which are on the surface at
time t will move off the surface at time t + Δt, sweeping out a volume given by
ΔV = Vn A Δt
~ ·n
where Vn = V ˆ is the component of the veleocity vector normal to the area.
V

n^
Vn V

swept volume n^

V Δt Vn Δt
V Δt
A A
The mass of fluid in this swept volume, which evidently passed through the area during the
Δt interval, is
Δm = ρ ΔV = ρ Vn A Δt
The mass flow is defined as the time rate of this mass passing though the area.
Δm
mass flow = m
˙ = lim = ρ Vn A
Δt→0 Δt
The mass flux is defined simply as mass flow per area.

mass flux = = ρVn
A

Mass Conservation Application


The conservation of mass principle can now be applied to the finite fixed control volume,
but now it must allow for the possibility of mass flow across the volume boundary.
d
(Mass in volume) = Mass flow into volume
dt
Using the previous relations we have
d ��� ��
~ ·n
ρ dV = − ρ V ˆ dA (1)
dt
where the negative sign is necessary because n̂ is defined to point outwards, so an inflow is
where −V~ ·n
ˆ is positive.

Using Gauss’s Theorem and bringing the time derivative inside the integral we have the

result ��� � �

∂ρ � �
~ dV = 0
+ ∇ · ρV
∂t

This relation must hold for any control volume whatsoever. If we place an infinitesimal
control volume at every point in the flow and apply the above equation, we can see that the
whole quantity in the brackets must be zero at every point. This results in the Continuity
Equation
∂ρ � �
+ ∇ · ρ V~ = 0 (2)
∂t
which is the embodiment of the Mass Conservation principle for fluid flow. The steady flow
version is: � �
∇ · ρV ~ = 0 (steady flow)
For low-speed flow, steady or unsteady, the density ρ is essentially constant, which gives the
very great simplification that the velocity vector field has zero divergence.
∇ · V~ = 0 (low-speed flow)

Control Volume Applications


Steady Surface-Integral Form
All of the above forms of the continuity equation are used in practice. The surface-integral
form (1) with the steady assumption,
��
~ ·n
ρV ˆ dA = 0 (3)

is particularly useful in many engineering applications.


Use of equation (3) first requires construction of the fixed control volume. The boundaries are
typically placed where the normal velocity and the density are known, so the surface integral
can be evaluated piecewise. Placing the boundary along a solid surface is particularly useful,
since here
~ ·n
V ˆ = 0 (at solid surface)
and so that part of the boundary doesn’t contribute to the integral. In the figure, we note
that
V~ ·n
ˆ dA = V cos θ dA = Vn dA = V dAs
~.
where dAs is the area of the surface element projected onto a plane normal to V

V Vn V

n^
θ
or

As
A A

The form of the integrand using dAs is most useful on a portion of the boundary which has
a constant value of ρV ~ (in magnitude and direction). The contribution of this boundary
portion to the overall integral (3) is then
�� ��
~ ·n
ρV ˆ dA = ρV dAs = ρV As ~)
(portion with constant ρV

3
V
V
boundary portion
with constant flow
As

~
Note that the boundary portion does not need to be flat for As to be well defined. Only ρV
must be a constant.

Channel Flow Application


Placing the control volume inside a pipe or channel of slowly-varying area, we now evaluate
equation (3).
��
~ ·n
ρV ˆ dA = −ρ1 V1 A1 + ρ2 V2 A2 = 0
ρ1 V1 A1 = ρ2 V2 A2
~ ·n
The negative sign for station 1 is due to V ˆ = −V1 at that location. The control volume
faces adjacent to walls do not contribute to the integral, since their normal vectors are
perpendicular to the local flow and therefore have V~ · n
ˆ = 0.

n^
V. n^ = 0 A2
A1
V2
n^ n^
V1

Placing plane 2, say, at any other location, gives the general result that

ρV A = constant

inside a channel. The product ρV A is also recognized as the constant channel mass flow.

Fluids – Lecture 7 Notes


1. Momentum Flow
2. Momentum Conservation
Reading: Anderson 2.5

Momentum Flow
Before we can apply the principle of momentum conservation to a fixed permeable control
volume, we must first examine the effect of flow through its surface. When material flows
through the surface, it carries not only mass, but momentum as well. The momentum flow
can be described as
−→ −→
momentum flow = (mass flow) × (momentum /mass)

where the mass flow was defined earlier, and the momentum/mass is simply the velocity
~ . Therefore
vector V
−→
~ = ρ V~ · n
ˆ A V~ = ρ Vn A V~
� �
momentum flow = m
˙ V

where Vn = V ~ ·n
ˆ as before. Note that while mass flow is a scalar, the momentum flow is a
~ . The momentum flux vector is defined simply
vector, and points in the same direction as V
as the momentum flow per area.
−→
~
momentum flux = ρ Vn V

V
n^ .
mV

.
A m

Momentum Conservation
Newton’s second law states that during a short time interval dt, the impulse of a force F~
applied to some affected mass, will produce a momentum change dP~a in that affected mass.
When applied to a fixed control volume, this principle becomes

F dP~a
= F~ (1)
dt
V
. dP~ ˙ ˙
. P(t) P out + P~ out − P~ in = F~ (2)
dt
P in

1
In the second equation (2), P~ is defined as the instantaneous momentum inside the control
volume. ���
P~ (t) ≡ ~ dV
ρV
˙
The P~ out is added because mass leaving the control volume carries away momentum provided
˙
by F~ , which P~ alone doesn’t account for. The P~ in is subtracted because mass flowing into
the control volume is incorrectly accounted in P~ , and hence must be discounted. Both terms
are evaluated by a surface integral of the momentum flux over the entire boundary.
˙ ˙
��
� �
P~ out − P~ in = ρ V~ ·n
ˆ V~ dA

~ ·n
The sign of V ˆ automatically accounts for both inlow and outflow.

Applied forces
The force F~ consists of two types.
Body forces. These act on fluid inside the volume. The most common example is the gravity
force, along the gravitational acceleration vector ~g .
���
F~gravity = ρ ~g dV

Surface forces. These act on the surface of the volume, and can be separated into pressure
and viscous forces. ��
F~pressure = −p n
ˆ dA

The viscous force is complicated to write out, and for now will simply be called F~viscous .

ρg −pn
Fviscous

−pn

Fviscous

Integral Momentum Equation


Substituting all the momentum, momentum flow, and force definitions into Newton’s second
law (2) gives the Integral Momentum Equation.

d
��� �� �� ���
~ dV + ρ V
~ ·n ~ dA = −p n ρ ~g dV + F~viscous
� �
ρV ˆ V ˆ dA + (3)
dt

2
Along with the Integral Mass Equation, this equation can be applied to solve many problems
involving finite control volumes.

Differential Momentum Equation


The pressure surface integral in equation (3) can be converted to a volume integral using the
Gradient Theorem. �� ���
p n̂ dA = ∇p dV
The momentum-flow surface integral is also similarly converted using Gauss’s Theorem. This
integral is a vector quantity, and for clarity the conversion is best done on each component
ˆ we have
separately. After substituting V~ = u ı̂ + v ̂ + w k,
�� ���
~ ·n u ı̂ + v ̂ + w kˆ dA = ı̂ ~ u dV
� � � � � �
ρ V ˆ ∇ · ρV
��� � �
+ ̂ ~ v dV
∇ · ρV
���
+ kˆ ∇ · ρ V~ w dV
� �

The x-component of the integral momentum equation (3) can now be written strictly in
terms of volume integrals.
��� � �
∂(ρu) �
~ + ∂p − ρgx − (Fx )viscous dV = 0

+ ∇ · ρuV (4)
∂t ∂x

This relation must hold for any control volume whatsoever. If we place an infinitesimal
control volume at every point in the flow and apply equation (4), we can see that the whole
quantity in the brackets must be zero at every point. This results in the x-Momentum
Equation
∂(ρu) �
~ = − ∂p + ρgx + (Fx )viscous

+ ∇ · ρuV (5)
∂t ∂x
and the y- and z-Momentum Equations follow by the same process.

∂(ρv) �
~ = − ∂p + ρgy + (Fy )viscous

+ ∇ · ρv V (6)
∂t ∂y

∂(ρw) �
~ = − ∂p + ρgz + (Fz )viscous

+ ∇ · ρw V (7)
∂t ∂z
These three equations are the embodiment of the Newton’s second law of motion, applied
at every point in the flowfield. The steady flow version has the ∂/∂t terms omitted.

Fluids – Lecture 8 Notes


1. Streamlines
2. Pathlines
3. Streaklines
Reading: Anderson 2.11

Three types of fluid element trajectories are defined: Streamlines, Pathlines, and Streaklines.
They are all equivalent for steady flows, but differ conceptually for unsteady flows.

Streamlines
Streamline equations
A streamline is defined as a line which is everywhere parallel to the local velocity vector
ˆ Define
~ (x, y, z, t) = u ı̂ + v ̂ + w k.
V
d~s = dx ı̂ + dy ̂ + dz k̂
~ , we
as an infinitesimal arc-length vector along the streamline. Since this is parallel to V
must have
~ = 0
d~s × V
(w dy − v dz) ı̂ + (u dz − w dx) ˆ + (v dx − u dy) ˆk = 0

Separately setting each component to zero gives three differential equations which define the
streamline. The three velocity components u, v, w, must be given as functions of x, y, z
before these equations can be integrated. To set the constants of integration, it is sufficient
to specify some point xo , yo, zo through which the streamline passes,
y y

ds V v
V
ds dy
dx u
xo yo z o
xo yo
x x

z 3−D streamline 2−D streamline

In 2-D we have dz = 0 and w = 0, and only the k̂ component of the equation above is
non-trivial. It can be written as an Ordinary Differential Equation for the streamline shape
y(x).
dy v
=
dx u
Again, u(x, y) and v(x, y) must be given to allow integration, and xo , yo must be given to
set the integration constants. In a numerical integration, xo , yo would serve as the initial
values.

1
Streamtubes
Consider a set of xo , yo , zo points arranged in a closed loop. The streamlines passing
through all these points form the surface of a streamtube. Because there is no flow across the
surface, each cross-section of the streamtube carries the same mass flow. So the streamtube
is equivalent to a channel flow embedded in the rest of the flowfield.
y y

xo yo z o xo yo
x x

z 3−D streamtube 2−D streamtube

In 2-D, a streamtube is defined by two streamlines passing through two specified xo , yo


points. The flow between these two streamlines carries the same mass flow/span at each
cross-section, and can be considered as a 2-D channel flow embedded in the rest of the
flowfield.

Pathlines
The pathline of a fluid element A is simply the path it takes through space as a function
of time. An example of a pathline is the trajectory taken by one puff of smoke which is
carried by the steady or unsteady wind. This path is fully described by the three position
functions xA (t), yA (t), zA (t), which can be computed by integrating the three velocity-field
components u(x, y, z, t), v(x, y, z, t), w(x, y, z, t) along the path. The integration is started
at time to , from the element’s initial position xo , yo , zo (e.g. the smoke release point), and
proceeds to some later time t.
� t � �
xA (t) = xo + u xA (τ ), yA (τ ), zA (τ ), τ dτ
to
� t � �
yA (t) = yo + v xA (τ ), yA (τ ), zA (τ ), τ dτ
to
� t � �
zA (t) = zo + w xA (τ ), yA (τ ), zA (τ ), τ dτ
to

The dummy variable of integration τ runs from to to t.

Streaklines
A streakline is associated with a particular point P in space which has the fluid moving past
it. All points which pass through this point are said to form the streakline of point P . An
example of a streakline is the continuous line of smoke emitted by a chimney at point P ,
which will have some curved shape if the wind has a time-varying direction.

2
Unlike a pathline, which involves the motion of only one fluid element A in time, a streakline
involves the motion of all the fluid elements along its length. Hence, the trajectory equations
for a pathline are applied to all the fluid elements defining the streakline.
The figure below illustrates streamlines, pathlines, and streaklines for the case of a smoke
being continuously emitted by a chimney at point P , in the presence of a shifting wind. One
particular smoke puff A is also identified. The figure corresponds to a snapshot when the
wind everywhere is along one particular direction.

s
streakline from point P a n eous
t e
(smoke line) tan lin
ins tream
s

A
P
instantaneous V streakline at
wind velocity successive times

pathline of fluid element A


(smoke puff)

In a steady flow, streamlines, pathlines, and streaklines all cooincide. In this example they
would all be marked by the smoke line.

Fluids – Lecture 9 Notes


1. Momentum-Integral Simplifications
2. Applications
Reading: Anderson 2.6

Simplifications
For steady flow, the momentum integral equation reduces to the following.
�� � � �� ���
~ ·n
ρ V ~ dA = −p n
ˆ V ˆ dA + ρ ~g dV + F~viscous (1)

Defining h as the height above ground, we note that ∇h is a unit vector which points up, so
that the gravity acceleration vector can be written as a gradient.

~g = −g ∇h

Using the Gradient Theorem, this then allows the gravity-force volume integral to be con-
verted to a surface integral, provided we make the additional assumption that ρ is nearly
constant throughout the flow.
��� ��� ��
ρ ~g dV = −ρ g ∇h dV = −ρgh n̂ dA

We can now combine the pressure and gravity contributions into one surface integral.
�� ��� ��
ˆ dA +
−p n ρ ~g dV → −(p + ρgh) n
ˆ dA

Defining a corrected pressure pc = p+ρgh, the Integral Momentum Equation finally becomes
�� � � ��
~ ·n
ρ V ~ dA = −pc n
ˆ V ˆ dA + F~viscous (2)

Aerodynamic analyses using (2) do not have to concern themselves with the effects of gravity,
since it does not appear explicitly in this equation. In particular, the velocity field V~ will
not be affected by gravity. Gravity enters the problem only in a secondary step, when the
true pressure field p is constructed from pc by adding the “tilting” bias −ρgh.
h h h h

= +
g

p pc ρg h
net aerodynamic aerostatic
(gravity neglected) (gravity alone)

1
For clarity, we will from now on refer to pc simply as p. The understanding is that the
forces and moments computed using this p will not include the contributions of buoyancy.
However, buoyancy is relatively easy to compute, and can be added as a secondary step
if appropriate. It should be noted that buoyancy is rarely significant in aerodynamic flow
situations, one exception being lighter-than-air vehicles like blimps.
One situation where gravity is crucial is a flow with a free surface, like water waves past a
ship. Here, gravity enters not in equation (2), but in the free-surface boundary condition.
Free-surface flows are not relevant to most aeronautical problems.

Applications
Basic Procedures
Application of the integral momentum equation (2) uses the same basic techniques as for
the integral continuity equation. Both can use the same control volume, and both demand
that the integrals are evaluated for the entire surface of the control volume. There are three
significant differences, however:
1) Momentum is a vector. Each of the three x, y, and z components of equation (2) is
independent, and must be treated separately.
2) A volume boundary lying along a streamline now has a nonzero −pn̂ contribution.
3) A volume boundary lying along a solid surface now has a nonzero −pn̂ contribution, and
possibly also a viscous shear ~τ contribution if viscous effects are deemed significant.

Effect of body in flow


Equation (2) assumes that there is no internal force acting on the fluid other than gravity.
If there is a solid body in the flow, it is therefore necessary to construct the control volume
such that the body is on the volume’s exterior. This is easily done by placing the body in
an indentation cdef g of the volume surface, as shown in the figure.

p1 = p p=p p2 = p
ay
streamline far aw
b
a .
dm
u1 u2
d c
e
f g

i
h
streamline far away

The surface integrals are now broken up into the individual pieces: the outermost part abhi,
the body surface def , and the “cut” surfaces cd and f g.
�� �� �� �� ��
( ) dA = ( ) dA + ( ) dA + ( ) dA + ( ) dA
abhi def cd fg

2
Surfaces cd and f g have the same flow variables, but opposite n̂ vectors, so that all their
contributions are equal and opposite, and hence cancel.
�� ��
( ) dA + ( ) dA = 0
cd fg

The resultant force/span R ~ ′


on the body is the integrated pressure and shear force distri-
bution on the body surface. Because the body surface and volume surface have opposite
~ ′ is precisely equal and opposite to all the def surface integrals for the
ˆ this R
normals n,
control volume.
��
( ) dA = −R ~ ′

def

p
R’ Forces on
airfoil
τ

d d
e e
f f

Forces on
τ control volume

p −R’

An intuitive explanation is that the aerodynamic force exerted by the fluid on the body is

exactly equal and opposite to the force which the body exerts on the fluid.

Collecting the remaining integrals produces the very general result

�� � � ��
ρ V~ · n̂ V
~ dA + ~′
p n̂ dA = −R (3)
abhi abhi

which gives the resultant force on the body only in terms of the surface integrals on the
outer surface of the volume.

Drag relations
To obtain the drag/span, we define the x direction to be aligned with V~∞ , and take the
x-component of equation (3).

�� � � ��
~ ·n
ρ V ˆ u dA + ˆ · ı̂ dA = −D ′

pn (4)
abhi abhi

The pressure on the outer surface is uniform and equal to p∞ , and hence the pressure inte-
gral is zero. The momentum-flux integral is zero on the top and bottom boundaries, since

these are defined to be along streamlines, and hence have zero momentum flux. Only the
momentum flux on the inflow and outflow planes remain.
�� ��
ρ2 u22 dA2 − ρ1 u21 dA1 = −D ′ (5)
bh ia

Using the concept of a streamtube, we see that planes 1 and 2 have the same mass flows,
with
ρ1 u1 dA1 = ρ2 u2 dA2 = dṁ
as shown in the figure. Therefore both the plane 1 and plane 2 integrals can be lumped into
one single plane 2 integral. � b
D ′
= ρ2 u2 (u1 − u2 ) dA2 (6)
h

Equation (6) gives the drag only in terms of the freestream speed u1 , and the downstream
wake “profiles” ρ2 and u2 . These downstream profiles can be measured in a wind tunnel,
and numerically integrated in equation (6) to obtain the drag.
A physical interpretation of (6) can be obtained by writing it as
� b

D = (u1 − u2 ) dṁ
h

This is simply the summation of the momentum flow “lost” by all the infinitesimal stream-
tubes in the flow. The lost momentum appears as drag on the body.

Fluids – Lecture 10 Notes


1. Substantial Derivative
2. Recast Governing Equations
Reading: Anderson 2.9, 2.10

Substantial Derivative
Sensed rates of change
The rate of change reported by a flow sensor clearly depends on the motion of the sensor.
For example, the pressure reported by a static-pressure sensor mounted on an airplane in
level flight shows zero rate of change. But a ground pressure sensor reports a nonzero rate
as the airplane rapidly flies by a few meters overhead. The figure illustrates the situation.
p
wing location at p1 (t)
t = to

p2 (t)

p2 (t) p1 (t)
t
to
Note that although the two sensors measure the same instantaneous static pressure at the
same point (at time t = to ), the measured time rates are different.
dp1 dp2
p1 (to ) = p2 (to ) but (to ) =
6 (to )
dt dt

Drifting sensor
We will now imagine a sensor drifting with a fluid element. In effect, the sensor follows the
element’s pathline coordinates xs (t), ys (t), zs (t), whose time rates of change are just the
local flow velocity components
dxs dys dzs
= u(xs , ys , zs , t) , = v(xs , ys , zs , t) , = w(xs , ys , zs , t)
dt dt dt
pressure ps
field

pathline

Dp dps
V Dt dt
ps (t)
t
sensor drifting with local velocity

Consider a flow field quantity to be observed by the drifting sensor, such as the static pressure
p(x, y, z, t). As the sensor moves through this field, the instantaneous pressure value reported
by the sensor is then simply

ps (t) = p (xs (t), ys (t), zs (t), t) (1)

1
This ps (t) signal is similar to p2 (t) in the example above, but not quite the same, since the
p2 sensor moves in a straight line relative to the wing rather than following a pathline like
the ps sensor.

Substantial derivative definition


The time rate of change of ps (t) can be computed from (1) using the chain rule.
dps ∂p dxs ∂p dys ∂p dzs ∂p
= + + +
dt ∂x dt ∂y dt ∂z dt ∂t
But since dxs /dt etc. are simply the local fluid velocity components, this rate can be ex-
pressed using the flowfield properties alone.
dps ∂p ∂p ∂p ∂p Dp
= + u + v + w ≡
dt ∂t ∂x ∂y ∂z Dt
The middle expression, conveniently denoted as Dp/Dt in shorthand, is called the substantial
derivative of p. Note that in order to compute Dp/Dt, we must know not only the p(x, y, z, t)
field, but also the velocity component fields u, v, w(x, y, z, t).
Although we used the pressure in this example, the substantial derivative can be computed
for any flowfield quantity (density, temperature, even velocity) which is a function of x, y, z, t.
D( ) ∂( ) ∂( ) ∂( ) ∂( ) ∂( ) ~ · ∇( )
≡ + u + v + w = + V
Dt ∂t ∂x ∂y ∂z ∂t

The rightmost compact D/Dt definition contains two terms. The first ∂/∂t term is called
the local derivative. The second V~ · ∇ term is called the convective derivative. In steady
flows, ∂/∂t = 0, and only the convective derivative contributes.

Recast Governing Equations


All the governing equations of fluid motion which were derived using control volume concepts
can be recast in terms of the substantial derivative. We will employ the following general
vector identity
∇ · (a~v ) = ~v · ∇a + a ∇·~v
which is valid for any scalar a and any vector ~v .

Continuity equation
Applying the above vector identity to the divergence form continuity equation gives
∂ρ �
~

+ ∇ · ρV = 0
∂t
∂ρ ~ · ∇ρ + ρ ∇· V
~ = 0
+ V
∂t
Dρ ~ = 0
+ ρ ∇· V (2)
Dt
The final result above is called the convective form of the continuity equation. A physical
interpretation can be made if it’s written as follows.

2
1 Dρ ~
− = ∇·V
ρ Dt
−fractional density rate = velocity divergence
or . . . fractional volume rate = velocity divergence

For a fluid element of given mass, the volume must vary as 1/density, which gives the second
interpretation above. Both interpretations are illustrated in the left figure below, where the
fluid element expands when it flows through a flowfield region where ∇ · V ~ > 0. In low
speed flows and in liquid flows the density is essentially constant, so that Dρ/Dt = 0 and
~ = 0.
by implication ∇· V

Momentum equation
The divergence form of the x-momentum equation is

∂(ρu) �
~ = − ∂p + ρgx + (Fx )viscous

+ ∇ · ρuV
∂t ∂x
Applying the vector identity again, and also cancelling some terms by use of the continuity
equation (2), produces the convective form of the momentum equation. The y- and z-
momentum equations are also derived the same way.
Du ∂p
ρ = − + ρgx + (Fx )viscous (3)
Dt ∂x
Dv ∂p
ρ = − + ρgy + (Fy )viscous (4)
Dt ∂y
Dw ∂p
ρ = − + ρgz + (Fz )viscous (5)
Dt ∂z
The Du/Dt etc. substantial derivatives are recognized as the acceleration components expe-
rienced by a fluid element. This leads to a simple physical interpretation or these equations
as Newton’s law applied to a fluid element of unit volume.

mass/volume × acceleration = total force/volume

The element’s mass/volume is simply the density ρ, and the total force/volume consists of
the buoyancy-like pressure gradient force, the gravity force, and the viscous force.

Δρ = D ρ Δ t ΔV = DV Δ t
Dt Dt

t + Δt V
t + Δt
t t
Δ .V p + ρ g +Fviscous
Δ−

expanding fluid element accelerating fluid element

Fluids – Lecture 11 Notes


1. Vorticity and Strain Rate
2. Circulation
Reading: Anderson 2.12, 2.13

Vorticity and Strain Rate


Fluid element behavior
When previously examining fluid motion, we considered only the changing position and
velocity of a fluid element. Now we will take a closer look, and examine the element’s
changing shape and orientation
.
Consider a moving fluid element which is initially rectangular, as shown in the figure. If
the velocity varies significantly across the extent of the element, its corners will not move in
unison, and the element will rotate and become distorted.
y

V(y+dy)

V(y)

x
z element at time t element at time t + Δt

In general, the edges of the element can undergo some combination of tilting and stretching.
For now we will consider only the tilting motions, because this has by far the greatest
implications for aerodynamics.
The figure below on the right shows two particular types of element-side tilting motions. If
adjacent sides tilt equally and in the same direction, we have pure rotation. If the adjacent
sides tilt equally and in opposite directions, we have pure shearing motion.
Both of these motions have strong implications. The absense of rotation will lead to a great
simplification in the equations of fluid motion. Shearing together with fluid viscosity produce
shear stresses, which are responsible for phenomena like drag and flow separation.
tilting

stretching

Rotation Shearing motion


(vorticity) (strain rate)
General edge movement Tilting edge movements

1
Side tilting analysis
Consider the 2-D element in the xy plane, at time t, and again at time t + Δt.

6u dy Δ t element at time t + Δt
6y
y element at time t
u Δt
−Δθ1

u + 6u dy Δθ2
B 6y
6v dx Δ t
6x
dy v Δt
v v + 6v dx
u 6x
A C
dx
x

Points A and B have an x-velocity which differs by ∂u/∂y dy. Over the time interval Δt
they will then have a difference in x-displacements equal to
∂u
ΔxB − ΔxA = dy Δt
∂y
and the associated angle change of side AB is
ΔxB − ΔxA ∂u
−Δθ1 = = Δt
dy ∂y
assuming small angles. A positive angle is defined counterclockwise. We now define a time
rate of change of this angle as follows.
dθ1 Δθ1 ∂u
= lim = −
dt Δt→0 Δt ∂y
Similar analysis of the angle rate of side AC gives
dθ2 ∂v
=
dt ∂x
Vorticity
The angular velocity of the element, about the z axis in this case, is defined as the average
angular velocity of sides AB and AC.
� � � �
1 dθ1 dθ2 1 ∂v ∂u
ωz = + = −
2 dt dt 2 ∂x ∂y
The same analysis in the xz and yz planes will give a 3-D element’s angular velocities ωy
and ωx . � � � �
1 ∂u ∂w 1 ∂w ∂v
ωy = − , ωx = −
2 ∂z ∂x 2 ∂y ∂z

These three angular velocities are the components of the angular velocity vector
.

~ω = ωx ı̂ + ωy ̂ + ωz kˆ

However, since 2~ω appears most frequently, it is convenient to define the vorticity vector ξ~
as simply twice ~ω .
� � � � � �
∂w ∂v ∂u ∂w ∂v ∂u
ξ~ = 2~ω = − ı̂ + − ̂ + − k̂
∂y ∂z ∂z ∂x ∂x ∂y

The components of the vorticity vector are recognized as the definitions of the curl of V~ ,
hence we have
ξ~ = ∇ × V~

Two types of flow can now be defined:


1) Rotational flow. Here ∇ × V~ 6= 0 at every point in the flow. The fluid elements move and
deform, and also rotate.
2) Irrotational flow. Here ∇ × V ~ = 0 at every point in the flow. The fluid elements move
and deform, but do not rotate.

The figure contrasts the two types of flow.

Rotational flows Irrotational flows

Strain rate
Using the same element-side angles Δθ1 , Δθ2 , we can define the strain of the fluid element.

strain = Δθ2 − Δθ1

This is the same as the strain used in solid mechanics. Here, we are more interested in the
strain rate, which is then simply
d(strain) dΔθ2 dΔθ1 ∂v ∂u
≡ εxy = − = +
dt dt dt ∂x ∂y
Similarly, the strain rates in the yz and zx planes are
∂w ∂v ∂u ∂w
εyz = + , εzx = +
∂y ∂z ∂z ∂x

Circulation

3
Consider a closed curve C in a velocity field as shown in the figure on the left. The instan-
taneous circulation around curve C is defined by

Γ ≡ − ~ · d~s
V
C

In 2-D, a line integral is counterclockwise by convention. But aerodynamicists like to define


circulation as positive clockwise, hence the minus sign.

dA
ds V ξ

C A

Circulation is closely linked to the vorticity in the flowfield. By Stokes’s Theorem,


� �� � ��
~ · d~s = − ~ ·n ξ~ · n

Γ ≡ − V ∇×V ˆ dA = − ˆ dA
C A A

where the integral is over the area A in the interior of C, shown in the above figure on the
ˆ is the unit vector normal to this area. In the 2-D xy plane, we have ξ~ = ξ kˆ and
right, and n
n ˆ
ˆ = k, in which case we have a simpler scalar form of the area integral.
��
Γ = − ξ dA (in 2-D)
A

From this integral one can interpret the vorticity as –circulation per area, or

ξ = −
dA

Irrotational flows, for which ξ = 0 by definition, therefore have Γ = 0 about any contour
inside the flowfield. Aerodynamic flows are typically of this type. The only restriction on
this general principle is that the contour must be reducible to a point while staying inside
the flowfield. A contours which contains a lifting airfoil, for example, is not reducible, and
will in general have a nonzero circulation.

Γ=0

Γ>0

Fluids – Lecture 12 Notes


1. Stream Function
2. Velocity Potential
Reading: Anderson 2.14, 2.15

Stream Function
Definition
~ as the partial derivatives
Consider defining the components of the 2-D mass flux vector ρV
¯ y):
of a scalar stream function, denoted by ψ(x,

∂ψ¯ ∂ψ¯
ρu = , ρv = −
∂y ∂x
For low speed flows, ρ is just a known constant, and it is more convenient to work with a
scaled stream function
ψ̄
ψ(x, y) =
ρ
~.
which then gives the components of the velocity vector V
∂ψ ∂ψ
u = , v = −
∂y ∂x

Example
Suppose we specify the constant-density streamfunction to be
� 1
ψ(x, y) = ln x2 + y 2 = ln(x2 + y 2)
2
which has a circular “funnel” shape as shown in the figure. The implied velocity components
are then
∂ψ y ∂ψ −x
u = = 2 , v = − = 2
∂y x +y 2 ∂x x + y2
which corresponds to a vortex flow around the origin.

vortex flow example


1
Streamline interpretation
The stream function can be interpreted in a number of ways. First we determine the differ-
ential of ψ¯ as follows.
∂ψ¯ ∂ψ¯
dψ¯ = dx + dy
∂x ∂y
dψ¯ = ρu dy − ρv dx

Now consider a line along which ψ¯ is some constant ψ¯1 .


¯ y) = ψ¯1
ψ(x,

Along this line, we can state that dψ¯ = dψ¯1 = d(constant) = 0, or


dy v
ρu dy − ρv dx = 0 → =
dx u
which is recognized as the equation for a streamline. Hence, lines of constant ψ(x, ¯ y) are
streamlines of the flow. Similarly, for the constant-density case, lines of constant ψ(x, y) are
streamlines of the flow. In the example above, the streamline defined by

ln x2 + y 2 = ψ1

can be seen to be a circle of radius exp(ψ1 ).


− −
y V ψ = ψ1

dψ = 0 v

dy

dx u

Mass flow interpretation


Consider two streamlines along which ψ¯ has constant values of ψ¯1 and ψ¯2 . The constant
mass flow between these streamlines can be computed by integrating the mass flux along
any curve AB spanning them. First we note the geometric relation along the curve,

n
ˆ dA = ı̂ dy − ˆdx

and the mass flow integration then proceeds as follows.


� B � B � B
m
˙ = ~ ·n
ρV ˆ dA = (ρu dy − ρv dx) = dψ¯ = ψ¯2 − ψ¯1
A A A

Hence, the mass flow between any two streamlines is given simply by the difference of their
stream function values.

2
− −
y ψ = ψ2
− − B
ψ + dψ V
V
dy dA
n^
− −

ψ n^ ψ = ψ1
A
x
dx

Continuity identity
¯ y). Computing the divergence
~ (x, y) specified by some ψ(x,
Consider the mass flux field ρV
of this field we have
¯ ¯ ∂ 2 ψ̄ ∂ 2 ψ̄
� � � �

~
� ∂(ρu) ∂(ρv) ∂ ∂ψ ∂ ∂ψ
∇ · ρV = + = − = − = 0
∂x ∂y ∂x ∂y ∂y ∂x ∂y ∂x ∂x ∂y
¯ y) will automatically satisfy the steady mass
so that any mass flux field specified via ψ(x,
continuity equation. In low speed flow, a similar computation shows that any velocity field
specified via ψ(x, y) will automatically satisfy
~ = 0
∇·V
which is the constant-density mass continuity equation. Because of these properties, using
the stream function to define the velocity field can give mathematical simplification in many
fluid flow problems, since the continuity equation then no longer needs to be addressed.

Velocity Potential
Definition
~ as the gradient of a scalar velocity
Consider defining the components of the velocity vector V
potential function, denoted by φ(x, y, z).

~ = ∇φ = ı̂ ∂φ + ˆ ∂φ + kˆ ∂φ
V
∂x ∂y ∂z
If we set the corresponding x, y, z components equal, we have the equivalent definitions
∂φ ∂φ ∂φ
u = , v = , w =
∂x ∂y ∂z

Example
For example, suppose we specify the potential function to be
� �
x
φ(x, y) = arctan
y
which has a corkscrew shape as shown in the figure. The implied velocity components are
then
∂φ y ∂φ −x
u = = 2 , v = = 2
∂x x +y 2 ∂y x + y2

3
φ

vortex flow example


which corresponds to a vortex flow around the origin. Note that this is exactly the same
velocity field as in the previous example using the stream function.

Irrotationality
If we attempt to compute the vorticity of the potential-derived velocity field by taking its
curl, we find that the vorticity vector is identically zero. For example, for the vorticity
x-component we find
∂w ∂v ∂ ∂φ ∂ ∂φ ∂2φ ∂2φ
ξx ≡ − = − = − = 0
∂y ∂z ∂y ∂z ∂z ∂y ∂y∂z ∂z∂y
and similarly we can also show that ξy = 0 and ξz = 0. This is of course just a manifestation
of the general vector identity curl(grad) = 0 . Hence, any velocity field defined in terms
of a velocity potential is automatically an irrotational flow . Often the synonymous term
potential flow is also used.

Directional Derivative
In many situations, only one particular component of the velocity is required. For example,
for computing the mass flow across a surface, we only require the normal velocity component.
This is typically computed via the dot product V~ · n.
ˆ In terms of the velocity potential, we
have
~ ·n ∂φ
V ˆ = ∇φ · n ˆ =
∂n
where the final partial derivative ∂φ/∂n is called the directional derivative of the potential
along the normal coordinate n. The figure illustrates the relations.

φ = φ4 V= φ
Δ
φ = φ3
φ = φ2 n
φ = φ1
V . n^ = 6φ
6n

In general, the component of the velocity along any direction can be obtained simply by
taking the directional derivative of the potential along that same direction.

4
Fluids – Lecture 13 Notes
1. Bernoulli Equation
2. Uses of Bernoulli Equation
Reading: Anderson 3.2, 3.3

Bernoulli Equation
Derivation – 1-D case
The 1-D momentum equation, which is Newton’s Second Law applied to fluid flow, is written
as follows.
∂u ∂u ∂p
ρ + ρu = − + ρgx + (Fx )viscous
∂t ∂x ∂x
We now make the following assumptions about the flow.
• Steady flow: ∂/∂t = 0
• Negligible gravity: ρgx ≃ 0
• Negligible viscous forces: (Fx )viscous ≃ 0
• Low-speed flow: ρ is constant
These reduce the momentum equation to the following simpler form, which can be immedi-
ately integrated.
du dp
ρu + = 0
dx dx
1 d(u2 ) dp
ρ + = 0
2 dx dx
1 2
ρ u + p = constant ≡ po
2
The final result is the one-dimensional Bernoulli Equation, which uniquely relates velocity
and pressure if the simplifying assumptions listed above are valid. The constant of integration
po is called the stagnation pressure, or equivalently the total pressure, and is typically set by
known upstream conditions.

Derivation – 2-D case


The 2-D momentum equations are
∂u ∂u ∂u ∂p
ρ + ρu + ρv = − + ρgx + (Fx )viscous
∂t ∂x ∂y ∂x
∂v ∂v ∂v ∂p
ρ + ρu + ρv = − + ρgy + (Fy )viscous
∂t ∂x ∂y ∂y
Making the same assumptions as before, these simplify to the following.
∂u ∂u ∂p
ρu + ρv + = 0 (1)
∂x ∂y ∂x
∂v ∂v ∂p
ρu + ρv + = 0 (2)
∂x ∂y ∂y

1
Before these can be integrated, we must first restrict ourselves only to flowfield variations
along a streamline. Consider an incremental distance ds along the streamline, with projec-
tions dx and dy in the two axis directions. The speed V likewise has projections u and
v.
y p + dp
u + du
v + dv V strea
mline
p
u
v v
dy

dx u

x
Along the streamline, we have
dy v
=
dx u
or
u dy = v dx (3)
We multiply the x-momentum equation (1) by dx, use relation (3) to replace v dx by u dy,
and combine the u-derivative terms into a du differential.
∂u ∂u ∂p
ρu dx + ρv dx + dx = 0
∂x ∂y ∂x
� �
∂u ∂u ∂p
ρu dx + dy + dx = 0
∂x ∂y ∂x
∂p
ρu du + dx = 0
∂x
1 � 2� ∂p
ρd u + dx = 0 (4)
2 ∂x
We multiply the y-momentum equation (2) by dy, and performing a similar manipulation,
we get
∂v ∂v ∂p
ρu dy + ρv dy + dy = 0
∂x ∂y ∂y
� �
∂v ∂v ∂p
ρv dx + dy + dy = 0
∂x ∂y ∂y
∂p
ρv dv + dy = 0
∂y
1 � � ∂p
ρ d v2 + dy = 0 (5)
2 ∂y
Finally, we add equations (4) and (5), giving
1 � 2 � ∂p ∂p
ρ d u + v2 + dx + dy = 0
2 ∂x ∂y
1 � 2 �
ρ d u + v 2 + dp = 0
2
2
which integrates into the general Bernoulli equation

1
ρ V 2 + p = constant ≡ po (along a streamline) (6)
2
where V 2 = u2 + v 2 is the square of the speed. For the 3-D case the final result is exactly
the same as equation (6), but now the w velocity component is nonzero, and hence V 2 =
u2 + v 2 + w 2 .

Irrotational Flow
Because of the assumptions used in the derivations above, in particular the streamline rela-
tion (3), the Bernoulli Equation (6) relates p and V only along any given streamline. Different
streamlines will in general have different po constants, so p and V cannot be directly related
between streamlines. For example, the simple shear flow on the left of the figure has parallel
flow with a linear u(y), and a uniform pressure p. Its po distribution is therefore parabolic
as shown. Hence, there is no unique correspondence between velocity and pressure in such
a flow.
y y
V V

po po

Rotational flow Irrotational flow

� = ∇φ and V 2 = |∇φ|2 , then po takes on the same


However, if the flow is irrotational, i.e. if V
value for all streamlines, and the Bernoulli Equation (6) becomes usable to relate p and V in
the entire irrotational flowfield. Fortunately, a flowfield is irrotational if the upstream flow
is irrotational (e.g. uniform), which is a very common occurance in aerodynamics. From the
uniform far upstream flow we can evaluate
1
po = p∞ + ρV∞2 ≡ po∞
2
and the Bernoulli equation (6) then takes the more general form.
1
ρ V 2 + p = po∞ (everywhere in an irrotational flow) (7)
2

Uses of Bernoulli Equation


Solving potential flows
Having the Bernoulli Equantion (7) in hand allows us to devise a relatively simple two-step
solution strategy for potential flows.
1. Determine the potential field φ(x, y, z) and resulting velocity field V� = ∇φ using the

governing equations.
2. Once the velocity field is known, insert it into the Bernoulli Equation to compute the
pressure field p(x, y, z).
This two-step process is simple enough to permit very economical aerodynamic solution
methods which give a great deal of physical insight into aerodynamic behavior. The alter-
native approaches which do not rely on Bernoulli Equation must solve for V � (x, y, z) and
p(x, y, z) simultaneously, which is a tremendously more difficult problem which can be ap-
proached only through brute force numerical computation.

Venturi flow
Another common application of the Bernoulli Equation is in a venturi , which is a flow tube
with a minimum cross-sectional area somewhere in the middle.

A1 A2

V1 V2
p
po
p1
p2
x
Assuming incompressible flow, with ρ constant, the mass conservation equation gives

A1 V1 = A2 V2 (8)

This relates V1 and V2 in terms of the geometric cross-sectional areas.


A1
V2 = V1
A2
Knowing the velocity relationship, the Bernoulli Equation then gives the pressure relation-
ship.
1 1
p1 + ρV12 = po = p2 + ρV22 (9)
2 2
Equations (8) and (9) together can be used to determine the inlet velocity V1 , knowing only
the pressure difference p1 − p2 and the geometric areas. By direct substution we have

2(p1 − p2 )


V1 = �
ρ [(A1 /A2 )2 − 1]

A venturi can therefore by used as an airspeed indicator, if some means of measuring the
pressure difference p1 − p2 is provided.

4
Fluids – Lecture 14 Notes
1. Helmholtz Equation
2. Incompressible Irrotational Flows
Reading: Anderson 3.7

Helmholtz Equation
Derivation (2-D)
If we neglect viscous forces, the x- and y-components of the 2-D momentum equation can
be written as follows.
∂u ∂u ∂u −1 ∂p
+ u + v = + gx (1)
∂t ∂x ∂y ρ ∂x
∂v ∂v ∂v −1 ∂p
+ u + v = + gy (2)
∂t ∂x ∂y ρ ∂y
We now take the curl of this momentum equation by performing the following operation.
∂ ∂
   
y-momentum (2) − x-momentum (1)
∂x ∂y

If we assume that ρ is constant (low speed flow), the two pressure derivative terms cancel.
Since the gravity components gx and gy are generally constant, these also disappear when
the curl’s derivatives are applied. Using the product rule on the lefthand side, the resulting
equation is
! ! !
∂ ∂v ∂u ∂ ∂v ∂u ∂ ∂v ∂u
− + u − + v −
∂t ∂x ∂y ∂x ∂x ∂y ∂y ∂x ∂y
!" #
∂v ∂u ∂u ∂v
+ − + = 0
∂x ∂y ∂x ∂y

We note that the quantity inside the parentheses is merely the z-component of the vorticity
ξ ≡ ∂v/∂x − ∂u/∂y, so the above equation can be more compactly written as
" #
∂ξ ∂ξ ∂ξ ∂u ∂v
+ u + v + ξ + = 0
∂t ∂x ∂y ∂x ∂y

We further note that the quantity in the brackets is the divergence of the velocity, which in
low speed flow must be zero because of mass conservation.
∂u ∂v ~ = 0
+ ≡ ∇·V (mass conservation equation)
∂x ∂y
This gives the following final result.
∂ξ ∂ξ ∂ξ
+ u + v = 0
∂t ∂x ∂y

or . . . = 0 (3)
Dt

1
Equation (3) is the 2-D form of the Helmholtz Equation, which governs the vorticity field
ξ(x, y, t) in inviscid flow.

Interpretation and Implications


Consider a microscopic sensor drifting with the flow (along a pathline) near an airfoil. The
sensor’s time-trace signal ξs (t) is the vorticity at the sensor’s instantaneous location. Equa-
tion (3) implies that this vorticity signal ξs (t) will have zero time rate of change, since we
know that
dξs Dξ
= = 0
dt Dt
Hence, the vorticity along a pathline must be constant.
ξs = constant
ξs

ξs(t) dξ s = 0
dt

t
Vorticity sensor moving along streamline Constant sensor output

Furthermore, this constant value must be determined far upstream of the airfoil where the
~∞ is uniform (either zero or some
streamline originates. If the freestream flow velocity V
constant), then
ξs = ∇ × V~∞ = 0
This is true for all streamlines which originate in the uniform upstream flow, so that the
entire flowfield must be irrotational, as shown in the figure.
~ = 0
ξ(x, y, t) ≡ ∇ × V (if upstream flow is uniform)
The one exception is that ξ 6= 0 for any streamline which is affected by viscous forces. For
these streamlines the Helmholtz equation (3) does not hold, since here the viscous forces are
not negligible, as was assumed at the outset.

ξ=0 here implies . . . ξ=0 everywhere downstream, except . . .


.
.
.

ξ=0
inside
boundary
layers

For 3-D flows, it is possible to derive a more general 3-D Helmholtz equation. From this
we can also conclude that 3-D flows which are initially uniform are irrotational downstream.
These 3-D derivations are considerably more cumbersome, and will not be attempted here.

2
Incompressible, Irrotational Flows
Governing Equations
The mass conservation equation for an incompressible flow states that the velocity field has
zero divergence.
∇·V~ = 0 (4)
The Helmholtz equation implies that an inviscid flow which is uniform upstream must be
irrotational, and can therefore be expressed in terms of a potential function.
~ = ∇φ
V

Substituting this into the divergence equation (4) gives

∇ · (∇φ) = ∇2 φ = 0 (5)

This is Laplace’s Equation. In Cartesian coordinates, with φ = φ(x, y, z), Laplace’s equation
is explicitly given by
∂2φ ∂2φ ∂2φ
∇2 φ = + + = 0
∂x2 ∂y 2 ∂z 2
In cylindrical coordinates, with φ = φ(r, θ, z), it has the form
!
2 1 ∂ ∂φ 1 ∂2φ ∂2φ
∇ φ = r + + = 0
r ∂r ∂r r 2 ∂θ2 ∂z 2

For 2-D problems, the stream function can be employed in lieu of the potential function.
The requirement that the flow be irrotational leads to
∂v ∂u ∂ ∂ψ ∂ ∂ψ
− = − − = 0
∂x ∂y ∂x ∂x ∂y ∂y
∂2ψ ∂2ψ
+ = 0 (6)
∂x2 ∂y 2
Hence, if the stream function is employed, it must also satisfy Laplace’s equation.

Superposition
Laplace’s equation is linear . If φ1 (x, y, z), φ2 (x, y, z) are valid solutions, then their sum

φ3 (x, y, z) = φ1 + φ2

is another valid solution. The corresponding velocities can therefore be obtained via vector
summation.
~3 (x, y, z) = ∇φ3 = ∇ (φ1 + φ2 ) = V
V ~1 + V ~2
This is the principle of superposition, which allows constructing complex flowfields from any
number of relatively simple components. The figure shows an example of two uniform flows
being superimposed into a third uniform flow. Stream functions can be superimposed in the
same manner. The pressure field in each case is obtained using Bernoulli’s equation.
1 1
p1 (x, y, z) = po − ρ |∇φ1 |2 , p2 (x, y, z) = po − ρ |∇φ2 |2 . . . etc
2 2
3
φ1 = x φ 2 = 0.5 y φ 3 = x + 0.5 y

+ =

Boundary Conditions
In order to solve Laplace’s equation, it is necessary to apply boundary conditions at all
boundaries of the flowfield. For most aerodynamic problems these fall into two categories.

Infinity Boundary Conditions


The flow far away from the body must approach the freestream velocity. Choosing the x
axis to be aligned with the freestream direction, we require
∂φ
u = = V∞ (at infinity)
∂x
If a stream function is used, the corresponding boundary condition would be
∂ψ
u = = V∞ (at infinity)
∂y

Wall Boundary Conditions


The flow adjacent to the wall is physically constrained to flow parallel, or tangent to the
wall. If the velocity vector is tangent, then its normal component must clearly be zero.

~ · n̂ = (∇φ) · n̂ = ∂φ = 0
V (on wall)
∂n
The boundary condition on the alternative stream function is
∂ψ
V~ · n̂ = = 0 (on wall)
∂s
where s is the arc length along the surface. This can also be specified as

ψ(s) = constant (on wall)

6φ = V
6x
n^
V
6φ = 0
6φ = V 6n
6x
6φ = V
6x
6φ = V
6x

4
Fluids – Lecture 15 Notes
1. Uniform flow, Sources, Sinks, Doublets
Reading: Anderson 3.9 – 3.12

Uniform Flow
Definition
A uniform flow consists of a velocity field where V~ = uı̂ + v̂ is a constant. In 2-D, this
velocity field is specified either by the freestream velocity components u∞ , v∞ , or by the
freestream speed V∞ and flow angle α.

u = u∞ = V∞ cos α
v = v∞ = V∞ sin α

Note also that V∞2 = u2∞ + v∞


2
. The corresponding potential and stream functions are

φ(x, y) = u∞ x + v∞ y = V∞ (x cos α + y sin α)


ψ(x, y) = u∞ y − v∞ x = V∞ (y cos α − x sin α)

V
v
α
u

Zero Divergence
A uniform flow is easily shown to have zero divergence

~ = ∂u∞ + ∂v∞ = 0
∇·V
∂x ∂y

since both u∞ and v∞ are constants. The equivalent statement is that φ(x, y) satisfies
Laplace’s equation.

∂ 2 (u∞ x + v∞ y) ∂ 2 (u∞ x + v∞ y)
∇2 φ = + = 0
∂x2 ∂y 2
Therefore, the uniform flow satisfies mass conservation.
Zero Curl
A uniform flow is also easily shown to be irrotational, or to have zero vorticity.
� �
~ ≡ ξ~ = ∂v∞ ∂u∞ ˆ
∇×V − k = 0
∂x ∂y

1
The equivalent irrotationality condition is that ψ(x, y) satisfies Laplace’s equation.

∂ 2 (u∞ y − v∞ x) ∂ 2 (u∞ y − v∞ x)
∇2 ψ = + = 0
∂x2 ∂y 2

Source and Sink


Definition
A 2-D source is most clearly specified in polar coordinates. The radial and tangential velocity
components are defined to be
Λ
Vr = , Vθ = 0
2π r
where Λ is a scaling constant called the source strength. The volume flow rate per unit span
V̇ ′ across a circle of radius r is computed as follows.


� 2π
~ ·n
� 2π � 2π Λ
V̇ = V ˆ dA = Vr r dθ = r dθ = Λ
0 0 0 2π r
Hence we see that the source strength Λ specifies the rate of volume flow issuing outward
from the source. If Λ is negative, the flow is inward, and the flow is called a sink .
y y


Vr
r
θ
x x

Cartesian representation
The cartesian velocity components of the source or sink are
Λ x
u(x, y) =
2π x + y 2
2

Λ y
v(x, y) =
2π x + y 2
2

and the corresponding potential and stream functions are as follows.


Λ � Λ
φ(x, y) = ln x2 + y 2 = ln r
2π 2π
Λ Λ
ψ(x, y) = arctan(y/x) = θ
2π 2π
2
It is easily verified that apart from the origin location (x, y) = (0, 0), these functions satisfy
∇2 φ = 0 and ∇2 ψ = 0, and hence represent physically-possible incompressible, irrotational
flows.

Singularities
The origin location (0, 0) is called a singular point of the source flow. As we approach this
point, the magnitude of the radial velocity tends to infinity as
1
Vr ∼
r
Hence the flow at the singular point is not physical, although this does not prevent us from
using the source to represent actual flows. We will simply need to ensure that the singular
point is located outside the flow region of interest.

Uniform Flow with Source


Two or more incompressible, irrotational flows can be combined by superposition, simply by
adding their velocity fields or their potential or stream function fields. Superposition of a
uniform flow in the x-direction and a source at the origin therefore has
Λ x
u(x, y) = + V∞
2π x + y 2
2

Λ y
v(x, y) =
2π x + y 2
2

Λ � Λ
or φ(x, y) = ln x2 + y 2 + V∞ x = ln r + V∞ r cos θ
2π 2π
Λ Λ
or ψ(x, y) = arctan(y/x) + V∞ y = θ + V∞ r sin θ
2π 2π
The figure shows the streamlines of the two basic flows, and also the combined flow.

The bullet-shaped heavy line on the combined flow corresponds to the dividing streamline,
which separates the fluid coming from the freestream and the fluid coming from the source.
If we replace the dividing streamline by a solid semi-infinite body of the same shape, the
flow about this body will be the same as the flow outside the dividing streamline in the
superimposed flow.

3
Uniform Flow with Source and Sink
We now superimpose a uniform flow in the x-direction, with a source located at (−ℓ/2, 0),
and a sink of equal and opposite strength located at (+ℓ/2, 0), plus a freestream.

Λ
ψ = (θ1 − θ2 ) + V∞ r sin θ

y

ψ
r1
r2

θ2
θ1
x

The figure on the right shows the streamlines of the combined flow. The heavy line again
indicates the dividing streamline, which traces out a Rankine oval . All the streamlines inside
the oval originate at the source on the left, and flow into the sink on the right. The net
volume outflow from the oval is zero. Again, the dividing streamline could be replaced by a
solid oval body of the same shape. The flow outside the oval then corresponds to the flow
about this body.

Doublet
Consider a source-sink pair with strengths ±Λ, located at (∓ℓ/2, 0). Now let the separation
distance ℓ approach zero, while simultaneously increasing the source and sink strengths such
that the product κ ≡ ℓΛ remains constant. The resulting flow is a doublet with strength κ.
κ κ sin θ
ψ = lim − Δθ = −
ℓ→0
κ=const.
2π ℓ 2π r

y y

ψ
r

Δθ
x x

4
A similar limiting process can be used to produce the doublet’s potential function.
κ cos θ
φ =
2π r

The streamline shapes of the doublet are obtained by setting


κ sin θ
ψ = − = c = constant
2π r
r = d sin θ
κ
where d = −
2πc
In polar coordinates this is the equation for circles of diameter d, centered on x, y = (0, ±d/2).

Nonlifting Flow over Circular Cylinder


Flowfield definition
We now superimpose a uniform flow with a doublet.
κ sin θ κ
� �
ψ = V∞ r sin θ − = V∞ r sin θ 1 −
2π r 2π V∞ r 2
� �
R2
or ψ = V∞ r sin θ 1 − 2
r
2
where R ≡ κ/(2πV∞ )

This corresponds to the flow about a circular cylinder of radius R.

The radial and tangential velocities can be obtained by differentiating the stream function
as follows.
� �
1 ∂ψ R2
Vr = = V∞ cos θ 1 − 2
r ∂θ r
� �
∂ψ R2
Vθ = − = −V∞ sin θ 1 + 2
∂r r

5
Surface velocities and pressures
On the surface of the cylinder where r = R, we have

Vr = 0
Vθ = −2V∞ sin θ

The maximum surface speed of 2V∞ occurs at θ = ±90◦ .

The surface pressure is then obtained using the Bernoulli equation

1 � 2 �
p(θ) = po − ρ Vr + Vθ2
2
Substituting Vr = 0 and Vθ (θ), and using the freestream value for the total pressure,
1 2
po = p∞ + ρV
2 ∞
gives the following surface pressure distribution.
1 2� �
p(θ) = p∞ − ρV∞ 1 − 4 sin2 θ
2
The corresponding pressure coefficient is also readily obtained.

p(θ) − p∞
Cp (θ) ≡ 1 = 1 − 4 sin2 θ
2
ρV 2

Fluids – Lecture 16 Notes


1. Vortex
2. Lifting flow about circular cylinder
Reading: Anderson 3.14 – 3.16

Vortex
Flowfield Definition
A vortex flow has the following radial and tangential velocity components
C
Vr = 0 , Vθ =
r
where C is a scaling constant. The circulation around any closed circuit is computed as
θ2 C
� � �
Γ ≡ − ~ · d~s = −
V Vθ r dθ = − r dθ = −C (θ2 − θ1 )
θ1 r
y y
V
r dθ
ds

x x

The integration range θ2 −θ1 = 2π if the circuit encircles the origin, but is zero otherwise.

−2πC , (circuit encircles origin)
Γ =
0 , (circuit doesn’t encircle origin)
y y
θ1 θ
2

θ1 θ
2

x x

In lieu of C, it is convenient to redefine the vortex velocity field directly in terms of the
circulation of any circuit which encloses the vortex origin.
Γ
Vθ = −
2π r
1
A positive Γ corresponds to clockwise flow, while a negative Γ corresponds to counterclock-
wise flow.

Cartesian representation
The cartesian velocity components of the vortex are
Γ y
u(x, y) =
2π x + y 2
2

Γ x
v(x, y) = −
2π x + y 2
2

and the corresponding potential and stream functions are as follows.


Γ Γ
φ(x, y) = − arctan(y/x) = − θ
2π 2π
Γ � Γ
ψ(x, y) = ln x2 + y 2 = ln r
2π 2π

Singularity
As with the source and doublet, the origin location (0, 0) is called a singular point of the
vortex flow. The magnitude of the tangential velocity tends to infinity as
1
Vθ ∼
r
Hence, the singular point must be located outside the flow region of interest.

Lifting Flow over Circular Cylinder


Flowfield definition
We now superimpose a uniform flow with a doublet and a vortex.
� �
R2 Γ
ψ = V∞ r sin θ 1 − 2 + ln r
r 2π

This corresponds to the flow about a circular cylinder of radius R as before, but now a
top/bottom assymetry is introduced by the vortex.

2
The radial and tangential velocities can be obtained by differentiating the stream function
as follows.
� �
1 ∂ψ R2
Vr = = V∞ cos θ 1 − 2
r ∂θ r
� �
∂ψ R2 Γ
Vθ = − = −V∞ sin θ 1 + 2 −
∂r r 2π r

Surface velocities and pressures


On the surface of the cylinder where r = R, we have

Vr = 0
Γ
Vθ = −2V∞ sin θ −
2π R
The corresponding surface pressure coefficient follows.
�2
V2 Γ 2Γ
� � �
Cp (θ) = 1 − = 1 − 4 sin2 θ − − sin θ (1)
V∞2 2π V∞ R π V∞ R

Forces
The resultant force/span is obtained by integrating the pressure forces over the surface of
the cylinder. � �
~ ′ ′ ′
R ≡ D ı̂ + L ̂ = −p n
ˆ dA = −(p − p∞ ) nˆ dA (2)
The constant p∞ which has been subtracted from p in the integrand does not change the
integrated result. This follows from the general identity

(constant) n̂ dA = 0

which holds for any closed body. y


ny n^

R dθ
nx
θ
x

Breaking up the resultant force/span (2) into separate x- and y-components, and dividing
by 21 ρV∞2 2R, we obtain expressions for the drag and lift coefficients.
1

cd = −Cp nx dA
2R
1

cℓ = −Cp ny dA
2R
3
Using the cylinder geometry relations
nx = cos θ , ny = sin θ , dA = R dθ
and substituting the Cp (θ) result (1) gives
� �2 �
1 2π 1 2π Γ 2Γ
� � � � �
2
cd = −Cp cos θ dθ = −1 + 4 sin θ + + sin θ cos θ dθ
2 0 2 0 2πV∞ R πV∞ R
� �2 �
1 � 2π 1 � 2π Γ 2Γ
� � �
cℓ = −Cp sin θ dθ = −1 + 4 sin2 θ + + sin θ sin θ dθ
2 0 2 0 2πV∞ R πV∞ R
After evaluating the integrals we obtain the final results.
cd = 0
Γ
cℓ =
V∞ R
The equivalent dimensional forms are
D ′
= 0 (3)

L = ρ V∞ Γ (4)
The result of zero drag (3) is known as d’Alembert’s Paradox , since it’s in direct conflict
with the observation that D ′ > 0 for all real bodies in a uniform flow. The explanation is of
course that viscosity has been neglected. The lift result (4) is known as the Kutta-Joukowski
Theorem, which will turn out to be valid for a 2-D body of any shape, not just for a circular
cylinder.

Real Cylinder Flows


Real viscous flow about a circular cylinder at large Reynolds numbers exhibits large amounts
of flow separation and drag. Normally the flow is symmetric between top and bottom, and
hence the lift is zero. However, if the cylinder has a rotational velocity, the separation is
pushed aft on the aft-going side and pushed forward on the forward-going side, resulting in
a flow assymetry. This assymetry has an associated nonzero circulation and a corresponding
lift. This phenomenon is known as the Magnus effect. Although the lift generated by
a rotating cylinder can match or exceed the lift achievable by a wing of similar size, the
cylinder is not a satisfactory lifting device because of its unavoidably large drag.
L’

boundary layer
separation

A rotating sphere also exhibits the Magnus effect, and here it has a strong influence on
many ball sports. The curveball pitch in baseball, the diving topspin volley in tennis, and
the sideways curving flight of a sliced golf ball are all due to the Magnus effect.

4
Fluids – Lecture 17 Notes
1. Flowfield prediction
2. Source Sheets
Reading: Anderson 3.17

Flowfield Prediction
Problem definition
The flowfield examples used so far were used to demonstrate the basic ideas behind the
method of superposition. We chose some combination of elementary flows (uniform flow,
sources, vortices, etc.), and then determined the resulting flowfield. The corresponding body
shape was determined from the shape of the dividing streamline. However, such an approach
is not practical for engineering applications, where we want to specify the body shape, rather
than have it as an outcome. The problem can therefore be stated as follows.
Given: Body shape Y (x), Freestream velocity V~∞
Determine: Superposition of suitable elementary flows which produce the velocity field
~ (x, y) about the body.
V
It turns out that sources, vortices, and doublets are not ideally suited to this task because
of their strong singularities. The constraint that these singularities must be inside the body
is difficult to meet, especially if the body is very slender. For this reason we now define
slightly more elaborate elementary flows which are smoother, and therefore better suited to
representing smooth bodies.

Source Sheets
Definition
Consider a sequence of flows where a single source of strength Λ is repeatedly subdivided
into smaller sources which are evenly distributed along a line segment of length ℓ. The limit
of this subdivision process is a source sheet of strength λ = Λ/ℓ.
Λ Λ Λ
Λ → 2 × → 4 × → 8 × ... λ
2 4 8

The units of Λ are length2 /time, while the units of λ are length/time (or velocity). Note
that the total source strength is not changed in this process.
The limiting process shown above has assumed that the sheet is straight, and that the
sources are uniformly subdivided and uniformly distributed along the sheet. Neither of
these assumptions are required. The subdivided sources can be distributed along any chosen
curve, in any chosen density. Hence, the source sheet can be curved, and its strength λ can
vary along the sheet.

1
Properties
Consider an infinitesimal length ds of the sheet. The infinitesimal source strength of that
piece is dΛ = λ ds, and the corresponding potential at some field point P at (x, y) is

dΛ λ
dφ = ln r = ln r ds
2π 2π
where r is the distance between point (x, y) and the point on the sheet.


dΛ = λ ds P
r (x,y)
λ(s)
ds
0 l
The potential of the entire sheet at point P is then obtained by integrating the infinitesimal
contributions all along the sheet.
� ℓ λ
φ(x, y) = ln r ds
0 2π
The shape of the sheet and the λ(s) distribution must be specified before this integral can
be evaluated. The velocity of the sheet is then obtained by taking the gradient of the result.
Note that to build up the entire flowfield, the integral must be evaluated for each point P in
the xy plane. In practice this is not necessary, since for engineering purposes the velocity is
required only at a small set of points, such as on the surface of a body to allow computation
of the pressure and the resultant force.
Consider now a simpler straight source sheet extending from (−ℓ/2, 0) to (ℓ/2, 0), with a
constant strength λ.
y
P
y
λ ds
x x
−l/2 s l/2
The potential and the velocity components at point P are given by

λ ℓ/2 �

φ(x, y) = ln (x − s)2 + y 2 ds (1)
2π −ℓ/2
∂φ λ ℓ/2 ∂ λ ℓ/2 x−s
� � � � �
u(x, y) = = ln (x − s) + y ds =
2 2 ds
∂x 2π −ℓ/2 ∂x 2π −ℓ/2 (x − s)2 + y 2
∂φ λ ℓ/2 ∂ λ ℓ/2 y
� � � � �
v(x, y) = = ln (x − s) + y ds =
2 2 ds
∂y 2π −ℓ/2 ∂y 2π −ℓ/2 (x − s)2 + y 2

These integrals can be evaluated, although the resulting expressions are cumbersome, and
not too important for our purposes here. The really interesting result is for the normal

velocity v(x, y) very close to the sheet, either just above at y = 0+ , or just below at y = 0− .
After the necessary integration, we find that
λ λ
v(x, 0+ ) = , v(x, 0− ) = − (flat, isolated source sheet)
2 2
λ V
V . n^ = λ/2
V . n^ = −λ/2

The normal velocity is then simply a constant λ/2 directed outward. But if any other
singularity or freestream is present, this additional velocity will be superimposed on each
side of the sheet. For example, if the sheet is immersed in a freestream, we will have
λ λ
v(x, 0+ ) = + v∞ , v(x, 0− ) = − + v∞
2 2
By taking the difference between the top and bottom points, any such additional velocity is
removed, giving the very general normal-velocity jump condition for any source sheet in any
situation.
~ · n̂ = λ
v(x, 0+ ) − v(x, 0− ) ≡ ΔV (2)

λ
V ΔV . n^ = λ
v
u

The advantage of using source sheets rather than sources to represent a flowfield is illustrated
in the figure below, which shows source sheets superimposed on a uniform flow to the right.
In each case the sheet’s strength λ is set so as to cancel the freestream’s component normal to
the sheet, giving a net zero normal flow. Hence, the sheet is ideally suited for representing a
solid surface of a body, since it can impose the physically necessary flow-tangency condition
~ · n̂ = 0 by suitably adjusting the sheet’s strength λ.
V

Modeling approach
The fact that the velocity field of a source sheet is smooth, without the troublesome 1/r

3
singularity of a point source, allows us to place some number of such sheets (or panels) end
to end on the surface of the body. We then determine the strengths λj of all the panels
j = 1, 2, . . . n such that the flow is tangent everywhere on the surface of the body. The
superposition also incidentally produces some flow inside the body, but this is not physical
and is simply ignored.

λj λ2
λ1
V

λn

uniform flow + n source panels resulting flowfield

This use of source sheets in this manner to represent a flow is the basis of the panel method ,
which is widely used to compute the flow about aerodynamic bodies of arbitrary shape. The
approach presented here is actually suitable only for non-lifting bodies such as fuselages.
For airfoils, wings, and other lifting bodies, vortices must be added in some form to enable
circulation to be represented. This modification will be treated later.

Solution technique
It is important to realize that each panel strength λj cannot be set independently of the
others. With more than one panel present, the velocity V ~ and hence the flow tangency
~ ·n
condition V ˆ = 0 at any point i on the surface is influenced not only by that panel’s λi ,
but also by the strengths λj of all the other panels. In tensor notation this can be written
as
~ ·n ~∞ · n
� �
V ˆ = Aij λj + V ˆi
i
where Aij is called the aerodynamic influence matrix , which can be computed once the
geometry of all the panels is decided.

P Vi

λj V
P
~ ·n
Requiring that V ˆ = 0 for each of the n panel midpoints gives the following.
~∞ · n
Aij λj = −V ˆi

This is a n × n linear system for the λj unknowns, which can be solved numerically using
matrix solution methods such as Gaussian elimination. With the λj determined, the velocity
and pressure (via Bernoulli) can then be computed at any point in the flowfield and on the
surface of the body. Forces are then computed by integrating the surface pressures. This
completes the aerodynamic analysis problem.

4
Fluids – Lecture 18 Notes
1. Prediction of Lift
2. Vortex Sheets
Reading: Anderson 3.17

Prediction of Lift
Limitations of Source Sheets
A point source has zero circulation about any circuit. Evaluating Γ using its definition we
have
r2 Λ Λ
� � �
Γ ≡ − ~ · d~s = −
V Vr dr = − dr = − (ln r2 − ln r1 ) = 0
r1 2πr 2π
which gives zero simply because r1 = r2 for any closed circuit, whether the origin is enclosed
or not. A source sheet, which effectively consists of infinitesimal sources, must have zero
circulation as well.
y Γ=0 Γ=0

r2 V
r1 ds

dr
x λ

Λ
This zero-circulation property of source sheets has severe consequences for flow represen-
tation. Any aerodynamic model consisting only of a freestream and superimposed source
sheets will have Γ = 0, and hence L′ = 0 as well. Hence, lifting flows cannot be represented
by source sheets alone.
This limitation is illustrated if we use source panels to model a flow expected to produce
lift, such as that on an airfoil at an angle of attack. Examination of the streamlines reveals
that the rear dividing streamline leaves the airfoil off one surface as shown in the figure. The
model also predicts an infinite velocity going around the sharp trailing edge.
source sheet model

Γ=0
L’ = 0

smooth flow−off
reality (Kutta condition)

Γ>0
L’ > 0

1
On real airfoils the flow always flows smoothly off the sharp trailing edge, with no large
local velocities. This smooth flow-off is known as the Kutta condition, and it must be
faithfully duplicated in any flow model which seeks to predict the lift correctly. Changing
the streamline pattern to force the flow smoothly off the trailing edge requires the addition
of circulation, which implies that vortices must be included in the flow representation in
some manner.

Vortex Sheets
Definition
Consider a sequence of flows where a single vortex of strength Γ is repeatedly subdivided
into smaller vortices which are evenly distributed along a line segment of length ℓ. The limit
of this subdivision process is a vortex sheet of strength γ = Γ/ℓ.
Γ Γ Γ
Γ → 2 × → 4 × → 8 × ... γ
2 4 8

Like with the source sheet strength λ, the units of γ are length/time (or velocity).

Properties
The analysis of the vortex sheet closely follows that of the source sheets. The potential of
the vortex sheet at point P is
ℓ γ

φ(x, y) = − θ ds
0 2π

P
d Γ= γ ds (x,y)
γ (s) θ

ds
0 l
For a straight vortex sheet extending from (−ℓ/2, 0) to (ℓ/2, 0), with a constant strength γ,
the potential and the velocity components at point P are given by
γ ℓ/2 y

φ(x, y) = − arctan ds
2π −ℓ/2 x−s
∂φ γ ℓ/2 ∂ y γ ℓ/2 y
� � � �
u(x, y) = = − arctan ds = ds
∂x 2π −ℓ/2 ∂x x−s 2π −ℓ/2 (x − s)2 + y 2
∂φ γ ℓ/2 ∂ y γ ℓ/2
� � � �
−x
v(x, y) = = − arctan ds = ds
∂y 2π −ℓ/2 ∂y x−s 2π −ℓ/2 (x − s)2 + y 2

2
As with the earlier source sheets, these integrals are cumbersome to evaluate in general. But
if we evaluate very close to the sheet, either just above at y = 0+ , or just below at y = 0− ,
the tangential velocity becomes very simple.
γ γ
u(x, 0+ ) = , u(x, 0− ) = − (flat, isolated vortex sheet)
2 2
V . s^ = γ / 2
γ
V

V . s^ = −γ / 2
The tangential velocity is then simply a constant γ/2 directed clockwise around the sheet.
By taking the difference between the upper and lower points at some x location,we obtain a
very general tangential-velocity jump condition for any vortex sheet in any situation.
~ · ŝ = γ
u(x, 0+ ) − u(x, 0− ) ≡ ΔV (1)

The figure shows the vortex sheet with a freestream superimposed. The surface velocity
vector pattern is very complicated, but the tangential velocity jump across the sheet is a
constant equal the γ at all points.
ΔV . s^ = γ
V γ
v
u
The advantage of using vortex sheets rather than vortices to represent a flowfield is illustrated
in the figure below. The left figure shows a vortex sheet superimposed on a uniform flow. The
vortex sheets smoothly deforms the flowfield in the manner required to impose circulation
and lift. The right figure shows the same airfoil as before, but now vortex panels have been
used instead of source panels to represent the flow. The nature of the vortex panels permits
the Kutta condition to be imposed, giving smooth flow off the trailing edge. The airfoil now
has the expected amount of lift.

3
Modeling approach and solution technique
As illustrated with the airfoil example, vortex panels provide an alternative way to model
the flow about a body, both for lifting and non-lifting bodies. The solution approach is
nearly the same as with source panels. The V~ · nˆ = 0 flow tangency condition is imposed for
each panel, but now the additional Kutta condition at the trailing edge is also imposed. The
resulting linear system is then solved for all the panel strengths γj . The surface velocities,
surface pressures, and overall forces can then be computed.

Types of panels used in practice


Vortex panels are by far the most widely used for 2-D problems, such as the flow about an
airfoil. Vortex panels can represent lifting or nonlifting flows equally well, so there is little
reason to use the more restrictive source panels.
For 3-D problems, however, vortex panels run into serious difficulties. The main problem
is that the sheet strength ~γ is now a vector lying in the sheet. The associated tangential
velocity of magnitude γ/2 is also in the sheet, and perpendicular to ~γ . In contrast, the source
panel strength λ is still a scalar in 3-D.

λ γ µ

3−D source sheet 3−D vortex sheet 3−D doublet sheet

Because ~γ is now a vector, it is not well suited for solution in a panel method. Instead, 3-D
panel methods employ doublet sheets, whose doublet strength µ (unrelated to viscosity) is a
scalar quantity, and can also represent lifting flows. The figure above conceptually shows a
doublet sheet. The axis of each infinitesimal doublet is oriented normal to the sheet, rather
than along the x-axis as in our previous examples.
Doublet sheets alone are sufficient to represent the flow about any lifting or nonlifting 3-D
body. However, most modern 3-D panel methods actually employ a combination of source
sheets and doublet sheets. Compared to using only doublet sheets, the source+doublet sheet
combination turns out to give the best accuracy for a given computational time. The details
of such combined methods are far beyond scope here.

Fluids – Lecture 19 Notes


1. Airfoils – Overview
Reading: Anderson 4.1–4.3

Airfoils – Overview
3-D wing context
The cross-sectional shape of a wing or other streamlined surface is called an airfoil . The
importance of this shape arises when we attempt to model or approximate the flow about
the 3-D surface as a collection of 2-D flows in the cross-sectional planes.
z y
L’
V
L Γ

3−D Wing
2−D Airfoil section flows

In each such 2-D plane, the airfoil is the aerodynamic body shape of interest. 2-D section
properties become functions of the spanwise coordinate y. Examples are L′ (y), Γ(y), etc.
Quantities of interest for the whole wing cn then be obtained by integrating over all the
sectional flows. For example,
� b/2
L = L′ (y) dy
−b/2

where b is the wing span. The airfoil shape is therefore an important item of interest, since
it is key in defining the individual section flows.
It must be stressed that the 2-D section flows are not completely independent, but rather
they influence each other’s effective angle of attack, or the apparent V~∞ direction in each 2-D
plane. Fortunately this complication does not prevent us from treating each 2-D plane as
though it was truly independent, since the angle of attack corrections can be added separately
later.

Nomenclature
The figure below shows the key terms used when dealing with airfoil geometry. The Mean
Camber Line is defined to lie halfway between the upper and lower surfaces.
Maximum
Thickness Maximum
Leading Edge Camber Mean Camber Line

Chord Line

Trailing Edge
Chord c

Aerodynamic Characterization
The freestream velocity vector V~∞ is defined by its magnitude V∞ = |V ~∞ |, and the angle of
attack α it makes with the airfoil’s chord line. The overall aerodynamic loads on the airfoil
are the resultant force/span vector R~ ′ and the moment/span M ′ , by convention taken about
the quarter-chord location. The resultant force is resolved into a lift force L′ and drag force
~∞ .
D ′ perpendicular and parallel to V

L’ R’

M’ D’

α V
c/4

The forces and moment are more conveniently nondimensionalized using the freestream dy-
namic pressure q∞ ≡ 12 ρ∞ V∞2 and the chord c, giving the lift, drag, and moment coefficients.

L′ D′ M′
cℓ ≡ , cd ≡ , cm ≡
q∞ c q∞ c q∞ c2
Dimensional analysis reveals that these will depend only on the angle of attack α, the
Reynolds number Re ≡ ρ∞ V∞ c/µ∞ , the Mach number M∞ ≡ V∞ /a∞ , and on the airfoil
shape.
cℓ , cd , cm = f ( α , Re , M∞ , airfoil shape )
For low speed flows, M∞ has virtually no effect. And for a given airfoil shape, we therefore
have

cℓ , cd , cm = f ( α , Re ) (low speed flow, given airfoil)

Typical cℓ (α) and cm (α curves for any given Re have a number of important features, as
shown in the figure. For moderate angles of attack, the cℓ (α) curve is nearly linear, and very
closely matches the one predicted by potential-flow theory (e.g. a panel method). At some
larger angle of attack, cℓ curve reaches a maximum value of cℓmax and then decreases. For
α’s beyond cℓmax the airfoil is said to be stalled , and exhibits varying amounts of separated
flow. An analogous situation occurs for large negative α’s.
Within the linear region, the cℓ (α) curve can be closely approximated with a linear fit.

cℓ (α) = a0 (α − αL=0) (away from stall)

Here, a0 is the lift-curve slope, and αL=0 is the zero-lift angle. These can be measured or
computed reasonably accurately with a potential-flow method.

2
cl potential−flow
cm
prediction

cl max

dc
a 0 = d αl

α α

αL=0

The moment coefficient cm (α), when defined about the quarter-chord point, is very nearly
constant away from stall. Again this is predicted well by potential-flow methods. Past stall,
the cm (α) curve deviates sharply from its constant value.
The drag coefficient cd can be plotted versus α, as shown in the figure on the left. However,
a more useful and more standard way is to plot cℓ vs cd , with α simply a dummy parameter
along the curve. This plot is called a drag polar, and is shown in the figure on the right.
cd cl

cl max

cl
low drag

cd
max
range

α
cdmin
α cd

cdmin

One reason for using the drag polar format is that when evaluating the aerodynamic per-
formance of an airfoil, the α values are not really relevant. All that matters is the drag
and how it compares to lift. The drag polar format compares these directly, and hence
summarizes the most important features of the airfoil’s drag characteristics in one plot. One
, or (cℓ /cd )max , which is where a line from the
such feature is the maximum lift-to-drag ratio
origin lies tangent to the polar curve. The cℓmax and cdmin values are also directly visible.
An aerodynamicist might also note the low-drag range of lift coefficients where the airfoil
naturally wants to operate.
It must be stressed that cd values are roughly 100 times smaller than typical maximum cℓ
values. Hence, the cd axis on a polar plot is greatly enlarged.

3
A sample polar plot and cℓ (α; Re) and cm (α; Re) curves for an actual sailplane airfoil are
shown below, for two different Reynolds numbers.

Fluids – Lecture 20 Notes


1. Airfoils – Detailed Look
Reading: http://www.av8n.com/how/htm/ Sections 3.1–3.3 (optional)

Airfoils – Detailed Look


Flow curvature and pressure gradients
The pressures acting on an airfoil are determined by the airfoil’s overall shape and the angle
of attack. However, it’s useful to examine how local pressures are approximately affected by
local geometry, and the surface curvature in particular.
Consider a location near the airfoil surface, ignoring the thin boundary layer. The local
flow speed is V , the local streamline radius of curvature is R. Another equivalent way to
define the curvature is κ = 1/R = dθ/ds, where θ is the inclination angle of the surface or
streamline, and s is the arc length. Positive κ is defined to be concave up as shown.
y

n
V s
θ v
R =κ −1 u
V
x
V θ
6p
6n local cartesian xy axes

To determine how the pressure varies normal to the surface, we align local xy axes tangent
and normal to the surface, and employ the y-momentum equation, with the viscous forces
neglected.
∂p ∂v ∂v
= −ρu − ρv
∂y ∂x ∂y
The Cartesian velocity components are related to the speed and the surface angle as follows.

u = V cos θ , v = V sin θ

At the origin where θ = 0, we then have


∂v ∂V ∂θ
v=0 , u=V , = sin θ + V cos θ = Vκ
∂x ∂x ∂x
Therefore, the normal pressure gradient along y = n is
∂p
= −ρ V 2 κ
∂n
This is the normal-momentum equation, sometimes also called the centrifugal formula. It
describes the physical requirement that there must be a transverse pressure gradient to force
fluid to flow along a curved streamline. It is valid for inviscid flows, at any Mach number.

Implications for surface pressures


Because of the influence on normal pressure gradients, changes in surface curvature are
expected to cause changes in surface pressure. If a common reference pressure exists away
from the wall, a concave corvature will produce a higher pressure towards the wall, while a
convex curvature will produce a lower pressure towards the wall. The figure below illustrates
the situation for a simple bump. The “+” and “-” symbols indicate expected changes in
pressure.
n n n

6p >0
6p < 0 6n 6p < 0
6n 6n

− p
+ p
+ p
The curvature/pressure-gradient relation also indicates the pressures which can be expected
on the surface of a body such as an airfoil. Examination of the streamline curvatures indicates
that for a symmetric airfoil at zero angle of attack, higher pressure is expected at the leading
and trailing edges, while lower pressure is expected along the sides.

− −
+ +
− −

For the same symmetrical airfoil at an angle of attack, the streamline pattern and the
pressures near the leading edge are now considerably different. The stagnation point moves
under the leading edge, and a strongly reduced pressure, called a “leading edge spike”, forms
at the leading edge point itself.

− − −
+ +
+

The actual surface pressure force vectors −Cp n̂ are shown for the NACA 0015 airfoil, at
α = 0◦ (cℓ = 0), and α = 10◦ (cℓ = 1.23). The Cp (x) distributions are also shown plotted.

These results were computed using a panel method, and therefore correspond to inviscid
irrotational incompressible flow. The drag is predicted to be zero (d’Alembert’s Paradox),
and the possibility of boundary layer separation is ignored. Despite these limitations, the
calculations are useful in that they simply reveal the intense pressure spike, which is known
to promote separation of the upper surface boundary layer, and thus degrades the airfoil’s
stall resistance. A corrective redesign of the airfoil would normally be undertaken if the
leading edge spike is deemed to be too strong.

Use of camber
An effective way to reduce the intensity of the leading edge spike is to add camber to the
airfoil. The NACA 4415 airfoil has the same 15% maximum thickness (relative to chord) as
the 0015, but it has a nonzero 4% maximum camber. The figures below show the cambered
NACA 4415 airfoil at the same same cℓ = 0 and cℓ = 1.23 as in the NACA 0015 case
(comparing at the same lift or cℓ is more meaningful than comparing at the same α).

3
The leading edge spike at the high angle of attack is indeed reduced considerably. The low
angle of attack case now has a “negative” spike on the bottom surface, but this is much
weaker and appears tolerable.
Although camber is seen to be attractive in the case above, too much camber is usually
detrimental. The figure below shows the NACA 8415 at the same cℓ = 0 and cℓ = 1.23 con-
ditions. The intense spike on the bottom surface shows the drawback of using the excessive
8% camber – the low cℓ (high speed) condition is likely to have excessive drag. Selection of
the ideal amount of camber is a major design choice for the airplane designer.

4
Fluids – Lecture 21 Notes
1. Airfoil Polar Relations

Airfoil Polar Relations


Wing Loading
For an aircraft in steady level flight, we have that lift equals weight.
1
L = ρ V 2 S CL = mg
2
1 mg 1
ρV2 =
2 S CL
The ratio mg/S is called the wing loading, and has the units of force/area, or pressure. The
lift coefficient CL can be interpreted here as the constant of proportionality between the
wing loading and the dynamic pressure. The airspeed is given explicitly by

mg 2
V =
S ρ CL

An important performance measure of many airplanes is their speed range, or their max/min
speed ratio. From the above equations we see that V is maximum when CL is minimum,
and vice versa.
1 2 mg 1
ρ Vmin =
2 S CLmax

We can therefore form the ratio �


Vmax CLmin
=
Vmin CLmax
so a large speed ratio requires a very large CL ratio. Efficient flight at any particular CL ,
whether large or small, requires that the corresponding CD is acceptably low. The acceptable
CL range can be discerned on a CD (CL ) drag polar of the aircraft. A major component of
this is the wing airfoil’s cd (cℓ ) drag polar.

Sample airfoil polars


The figures show two airfoils for RC aircraft.

Dragonfly. This airfoil is used on the Dragonfly light electric sport aircraft.

AG44ct. This airfoil is used on high-performance composite RC sailplanes.

The most striking difference in the two airfoils is the camber:

Dragonfly: 7.3% camber

AG44ct(-2) 1.9% camber

AG44ct(+4) 3.4% camber

The main consequence is for the minimum flyable CL . The Dragonfly airfoil cannot operate
much below CLmin = 0.5 without incurring a massive drag increase (due to the bottom surface
stalling). The AG44ct(-2) in contrast can operate very near zero CL , with CLmin = 0.05 being

a practical lower limit. The maximum usable lift coefficients are roughly CLmax = 1.3 for the
Dragonfly, and CLmax = 0.85 for the RC sailplane with maximum camber flap deployed.
Both the Dragonfly and a typical RC sailplane have comparable wing loadings:
mg
≃ 15 Pa
S
The corresponding minimum and maximum speeds for the Dragonfly are

Vmin = 4.3 m/s = 9.7 mph (Dragonfly)


Vmax = 7.0 m/s = 15.6 mph (Dragonfly)

For the RC saiplane they are

Vmin = 5.3 m/s = 12.0 mph (RC sailplane)


Vmax = 22 m/s = 49.3 mph (RC sailplane)

Note the much larger speed range of the RC sailplane, which is important for fast ranging in
search of thermals, possibly against the wind. The sport Dragonfly has no such performance
requirement, and its narrow speed range is not a serious handicap.

You might also like