Anda di halaman 1dari 257

13.

021 – Marine Hydrodynamics, Fall 2004


Lecture 1

13.021 – Marine Hydrodynamics


Lecture 1

Introduction
Marine Hydrodynamics is the branch of Fluid Mechanics that studies the motion of incom­
pressible fluids (liquids) and the forces acting on solid bodies immersed in them.

Marine hydrodynamics is a large and diverse subject and only a limited number of topics
can be covered in an introductory course. The topics that will be covered throughout the
semester include:

• Model testing and similitude

• Interaction between bodies and ideal fluids

• Viscosity and boundary layers

• Effect of waves on resistance and ship motion

Why study Marine Hydrodynamics?


Studying Marine Hydrodynamics provides a greater understanding of a wide range of phe­
nomena of considerable complexity involving fluids. Another benefit is that it allows pre­
dictions to be made in many areas of practical importance.

Fluids vs. Solids


In brief, Fluid Mechanics studies the kinematics and dynamics of a group of particles with­
out having to study each particle separately.

Most of us have taken some courses on solids or related to solids. Even those who haven’t
can get an intuitive feeling about some physical properties of a solid. Thus a comparison
of solids and fluids will give some guidelines as to which properties can be translated to
fluids and on what terms.

Similarities

1. Fundamental laws of mechanics apply to both fluids and solids

- Conservation of mass
- Conservation of momentum (Newton’s law of motion)
- Conservation of energy (First law of thermodynamics)

2. Continuum hypothesis is used for both fluids and solids


measured property

“particle” size
local value

length scale

variations due to variations due to


molecular fluctuations varying flow
O(10-10 – 10-8)m

e.g. The smallest measurement scales are in the order of M ∼ 10−5 m → VM ∼ 10−15 m3 .
This corresponds to ∼ 3 × 1010 molecules of air in STP or ∼ 1013 molecules of water.

Differences

1. Shape

- Solids have definite shape


- Fluids have no preferred shape

2. Constitutive laws

Constitutive laws are empirical formulas that relate certain unknown variables.

The constitutive laws used in 13.021 relate:

dynamics (force, stress ...) to kinematics (position, displacement, velocity ...)

Different constitutive laws are used for solids and different for fluids.

- For solids Hooke’s law is used to relate stress and strain

�stress
�� � = f (strain)
� �� �
force/area relative displacement/length
� �
N/m2 [m/m]

•• For solid mechanics ‘statics’ is a dominant aspect

- For fluids stress is related to rate of strain

�stress
�� � = f (rate of strain)
� �� �
force/area velocity gradient
� �
N/m2 [1/s]

•• Fluids at rest cannot sustain shear force. Fluids have to be moving to be non-trivial.

The branch of Fluid Mechanics that studies fluids at rest is referred to as ‘Hydrostatics’.

(Archimedes, c 200 BC)

Hydrostatics study the trivial case where no stresses due to fluid motion exist.

Sometimes distinction between liquids and solids is not a sharp one(honey, jelly, paint, . . . ).
Fortunately most common fluids, such as air and water are very close to ”ideal” fluids.

Liquids vs. Gasses


Liquids and gasses are two categories of fluids.

A fluid is ‘a body whose particles move easily among themselves. Fluid is a generic term,
including liquids and gasses as species. Water, air, and steam are fluids.’ [1] .

A liquid is ‘Being in such a state that the component parts move freely among themselves,
but do not tend to separate from each other as the particles of gases and vapors do; neither
solid nor aeriform.’[1]

A gas is ‘The state of matter distinguished from the solid and liquid states by relatively low
density and viscosity, relatively great expansion and contraction with changes in pressure
and temperature, the ability to diffuse readily, and the spontaneous tendency to become dis­
tributed uniformly throughout any container.’[2]

In brief, a liquid is generally incompressible and does not fill a volume by expanding into
it while on the other hand, a gas is compressible and expands to fill any volume containing it.

The science that studies the dynamics of liquids is referred to as ‘Hydrodynamics’, while
the science that studies the dynamics of gasses is referred to as ‘Aerodynamics’.

The main difference between the study of ‘Hydrodynamics’ and the study of ‘Aerodynam­
ics’ is the property of incompressibility. In general hydrodynamic flows are treated as
incompressible while aerodynamic flows are treated as compressible.

[1] Webster Dictionary

[2] American Heritage Dictionary

4
Why is a liquid flow incompressible?

It can be shown that the ratio of the characteristic fluid velocities U in a flow to the
speed of sound C in the medium gives a measure of compressibility of the medium for that
particular flow. This ratio is called the Mach number M . Although the speed of sound
in water is of comparable magnitude to the speed of sound in air, the characteristic fluid
velocities in water are significantly smaller. Thus in the case of water, the Mach number
is very small, indicating that water is virtually incompressible.

U : Characteristic fluid flow velocity


C : Speed of sound in the medium
U
M≡ : Mach number
C
The average speed of sound in air and water is:

Cair ∼ 300m/s = 984f t/sec = 583knots


Cwater ∼ 1200m/s = 3, 937f t/sec = 2, 333knots
Cwater
Therefore the average ratio of the speed of sound in water to air is Cair
∼ 4. Further on,
ρwater 1kg/m3 3
because the average water to air density ratio is ρair
∼ 103 kg/m3
= 10 , it is ‘harder’ to
move in water and therefore, typically, it is:

Uwater <<Uair
giving thus typical values of Mach numbers in the order of:

Mair ∼ O(1) ⇒ COMPRESSIBLE flow


Mwater <<1 ⇒ INCOMPRESSIBLE flow

Note: An incompressible flow does not mean constant density.

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 2

13.021 – Marine Hydrodynamics


Lecture 2

Chapter 1 - Basic Equations


1.1 Description of a Flow
To define a flow we use either the ‘Lagrangian’ description or the ‘Eulerian’ description.
• Lagrangian description: Picture a fluid flow where each fluid particle caries its own
properties such as density, momentum, etc. As the particle advances its properties
may change in time. The procedure of describing the entire flow by recording the
detailed histories of each fluid particle is the Lagrangian description. A neutrally
buoyant probe is an example of a Lagrangian measuring device.

The particle properties density, velocity, pressure, . . . can be mathematically repre­


sented as follows: ρp (t), vp (t), pp (t), . . .

The Lagrangian description is simple to understand: conservation of mass and New­


ton’s laws apply directly to each fluid particle . However, it is computationally
expensive to keep track of the trajectories of all the fluid particles in a flow and
therefore the Lagrangian description is used only in some numerical simulations.

r
υ p (t )
p

Lagrangian description; snapshot

• Eulerian description: Rather than following each fluid particle we can record the
evolution of the flow properties at every point in space as time varies. This is the
Eulerian description. It is a field description. A probe fixed in space is an example
of an Eulerian measuring device.

This means that the flow properties at a specified location depend on the location
and on time. For example, the density, velocity, pressure, . . . can be mathematically
represented as follows: v (x, t), p(x, t), ρ(x, t), . . .

The aforementioned locations are described in coordinate systems. In 13.021 we use


the cartesian, cylindrical and spherical coordinate systems.

The Eulerian description is harder to understand: how do we apply the conservation


laws? However, it turns out that it is mathematically simpler to apply. For this
reason, in Fluid Mechanics we use mainly the Eulerian description.

y
r r
υ ( x, t )
r
( x, t )
x

Eulerian description; Cartesian grid

1.2 Flow visualization - Flow lines


• Streamline: A line everywhere tangent to the fluid velocity v at a given instant (flow
snapshot). It is a strictly Eulerian concept.

• Streakline: Instantaneous locus of all fluid particles that have passed a given point
(snapshot of certain fluid particles).

• Pathline: The trajectory of a given particle P in time. The photograph analogy would
be a long time exposure of a marked particle. It is a strictly Lagrangian concept.

Can you tell whether any of the following figures ( [1] Van Dyke, An Album of Fluid Motion
1982 (p.52, 100)) show streamlines/streaklines/pathlines?

1.3 Some Quantities of Interest


• Einstein Notation
– Range convention: Whenever a subscript appears only once in a term, the sub­
script takes all possible values. E.g. in 3D space:
xi (i = 1, 2, 3) → x1 , x2 , x3
– Summation convention: Whenever a subscript appears twice in the same term
the repeated index is summed over the index parameter space. E.g. in 3D space:
ai bi = a1 b1 + a2 b2 + a3 b3 (i = 1, 2, 3)
Non repeated subscripts remain fixed during the summation. E.g. in 3D space
ai = xij n̂j denotes three equations, one for each i = 1, 2, 3 and j is the dummy
index.

Note 1: To avoid confusion between fixed and repeated indices or different re­
peated indices, etc, no index can be repeated more than twice.

Note 2: Number of free indices shows how many quantities are represented by
a single term.

Note 3: If the equation looks like this: (ui ) (x̂i ) , the indices are not summed.

– Comma convention: A subscript comma followed by an index indicates partial


differentiation with respect to each coordinate. Summation and range conven­
tions apply to indices following a comma as well. E.g. in 3D space:
∂ui ∂u1 ∂u2 ∂u3
ui,i = = + +
∂xi ∂x1 ∂x2 ∂x3
• Scalars, Vectors and Tensors

Scalars Vectors (ai xi ) Tensors (aij )


magnitude magnitude magnitude
direction direction
orientation
density ρ (x, t) velocity v (x, t) /momentum momentum flux
pressure p (x, t) mass flux stress τij (x, t)

1.4 Concept and Consequences of Continuous Flow


For a fluid flow to be continuous, we require that the velocity v (x, t) be a finite and con­
tinuous function of x and t.
i.e. ∇ · v and ∂
v
∂t
are finite but not necessarily continuous.
∂v
Since ∇ · v and ∂t < ∞, there is no infinite acceleration i.e. no infinite forces , which is
physically consistent.

1.4.1 Consequences of Continuous Flow

• Material volume remains material. No segment of fluid can be joined or broken apart.

• Material surface remains material. The interface between two material volumes al­
ways exists.

• Material line remains material. The interface of two material surfaces always exists.

Material
surface

fluid a

fluid b

• Material neighbors remain neighbors. To prove this mathematically, we must prove


that, given two particles, the distance between them at time t is small, and the
distance between them at time t + δt is still small.

Assumptions At time t, assume a continuous flow (∇ · v , ∂ v


∂t
<< ∞) with fluid velocity
v (x, t). Two arbitrary particles are located at x and x + δx(t), respectively.

Result If δx(t) ≡ δx → 0 then δx(t + T ) → 0, for all subsequent times t + T .


v r v
Proof v x + {υ (x )δt}
x

v
v
δ x δx(t + δt)

v v
v v x +δ x +
x +δ x
{(υr (xv ) + δ xv ⋅∇υr (xv ) )δt}
After a small time δt:

• The particle initially located at x will have travelled a distance v (x)δt and at time
t + δt will be located at x + {v (x)δt}.

• The particle initially located at x+δx will have travelled a distance (v (x) + δx · ∇v (x)) δt.
(Show this using Taylor Series Expansion about(x, t)). Therefore after a small time
δt this particle will be located at x + δx + {(v (x) + δx · ∇v (x))δt}.

• The difference in position δx(t + δt) between the two particles after a small time δt
will be:

δx(t + δt) = x + δx + {(v (x) + δx · ∇v (x))δt} − (x + {v (x)δt})
⇒ δx(t + δt) = δx + (δx · ∇v (x))δt ∝ δx

Therefore δx(t + δt) ∝ δx because ∇v is finite (from continuous flow assumption).
Thus, if δx → 0 , then δx(t + δt) → 0. In fact, for any subsequent time t + T :
� t+T
δx(t + T ) ∝ δx + δx · ∇v dt ∝ δx,
t

and δx(t + T ) → 0 as δx → 0. In other words the particles will never be an infinite distance
apart. Thus, if the flow is continuous two particles that are neighbors will always remain
neighbors.

6
1.5 Material/Substantial/Total Time Derivative: D/Dt
A material derivative is the time derivative – rate of change – of a property following a
fluid particle ‘p’. The material derivative is a Lagrangian concept.

By expressing the material derivative in terms of Eulerian quantities we will be able to


apply the conservation laws in the Eulerian reference frame.

Consider an Eulerian quantity f (x, t). The time rate of change of f as experienced by a
particle ‘p’ travelling with velocity −

vp is the substantial derivative of f and is given by:

r r
f (x p + υ pδt , t + δt )
r r
Df (−

xp (t), t) f (−

xp + −

vp δt, t + δt) − f (−

xp , t) r x p (t ) + υ pδt
= lim (1) f ( x p ,t )
Dt δt→0 δt particle r
x p (t )

Performing a Taylor Series Expansion about (− →


xp , t) and taking into account that → −
vp δt = δx,
we obtain:
→+−
− → ∂f (x, t)
f (x p vp δt, t + δt) = f (x, t) + δt + δx · ∇f (x, t) + O(δ 2 )(higher order terms)(2)
∂t
From Eq.(1, 2) we see that the substantial derivative of f as experienced by a particle
travelling with →

vp is given by:
Df ∂f −
= +→
vp · ∇f
Dt ∂t
The generalized notation:
D ∂
≡ +−→vp · ∇
Dt
���� �∂t �� �
Lagrangian Eulerian

 (x, t) as experienced by a fluid par­


Example 1: Material derivative of a fluid property G
ticle.

Let ‘p’ denote a fluid particle. A fluid particle is always travelling with the local fluid
 (x, t) as experienced
velocity vp (t) = v (xp , t). The material derivative of a fluid property G
by this fluid particle is given by:

DG 
∂G
= + �v ·��
∇G
Dt ∂t �
���� ����
Convective
Lagragian Eulerian
rate of change
rate of change rate of change

Example 2: Material derivative of the fluid velocity v (x, t) as experienced by a fluid par­
ticle. This is the Lagrangian acceleration of a particle and is the acceleration that appears
in Newton’s laws. It is therefore evident that its Eulerian representation will be used in
the Eulerian reference frame.

Let ‘p’ denote a fluid particle. A fluid particle is always travelling with the local fluid
velocity vp (t) = v (xp , t). The Lagrangian acceleration DvDt
(
x,t)
as experienced by this fluid
particle is given by:

Dv ∂v
= + �v ·��
∇v�
Dt
���� ∂t
����
Convective
Lagragian Eulerian
acceleration
acceleration acceleration

1.6 Difference Between Lagrangian Time Derivative and


Eulerian Time Derivative

Example 1: Consider an Eulerian quantity, temperature, in a room at points A and B


where the temperature is different at each point.

∂T Point B: 1o
Point A: 10o Point C:
∂t

∂T
At a fixed in space point C, the temperature rate of change is ∂t
which is an Eulerian time
derivative.

Example 2: Consider the same example as above: an Eulerian quantity, temperature, in


a room at points A and B where the temperature varies with time.

DT ∂T v
= + υfly⋅ ∇ T
Point A: 10o Dt ∂t Point B: 1 o

Following a fly from point A to B, the Lagrangian time derivative would need to include
the temperature gradient as both time and position changes: DT
Dt
= ∂T
∂t
+ vf ly · ∇T


1.6.1 Concept of a Steady Flow ( ∂t ≡ 0)
A steady flow is a strictly Eulerian concept.

Assume a steady flow where the flow is observed from a fixed position. This is like watching

∂ D
from a river bank, i.e. ∂t = 0 . Be careful not to confuse this with Dt which is more like
D
following a twig in the water. Note that Dt = 0 does not mean steady since the flow could
speed up at some points and slow down at others.


=0
∂t

1.6.2 Concept of an Incompressible Flow ( Dρ Dt


≡ 0)
An incompressible flow is a strictly Lagrangian concept.

Assume a flow where the density of each fluid particle is constant in time. Be careful not
to confuse this with ∂ρ
∂t
= 0, which means that the density at a particular point in the flow
is constant and would allow particles to change density as they flow from point to point.
Also, do not confuse this with ρ = const , which for example does not allow a flow of two
incompressible fluids.

10

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 3

13.021 – Marine Hydrodynamics


Lecture 3

1.7 Stress Tensor


1.7.1 Stress Tensor τij
The stress (force per unit area) at a point in a fluid needs nine components to be completely
specified, since each component of the stress must be defined not only by the direction in
which it acts but also the orientation of the surface upon which it is acting.

The first index i specifies the direction in which the stress component acts, and the second
index j identifies the orientation of the surface upon which it is acting. Therefore, the ith
component of the force acting on a surface whose outward normal points in the j th direction
is τij .

X2

22

12
32
21
23
11
13 31
33 X1

X3

Figure 1: Shear stresses on an infinitesimal cube whose surfaces are parallel to the coordinate
system.

X2
2

P n̂
A1

1
A3
X1
Q
3
R
area A0
A2
X3

Figure 2: Infinitesimal body with surface PQR that is not perpendicular to any of the Cartesian
axis.

Consider an infinitesimal body at rest with a surface PQR that is not perpendicular to any
of the Cartesian axis. The unit normal vector to the surface PQR is n̂ = n1 x̂1 +n2 x̂2 +n3 x̂3 .
The area of the surface = A0 , and the area of each surface perpendicular to Xi is Ai = A0 ni ,
for i = 1, 2, 3.

Newton’s law: Fi = (volume force)i for i = 1, 2, 3
on all 4 faces

Note: If δ is the typical dimension of the body : surface forces ∼ δ 2


: volume forces ∼ δ 3

An example of surface forces is the shear force and an example of volumetric forces is the
gravity force. At equilibrium, the surface forces and volumetric forces are in balance. As
the body gets smaller, the mass of the body goes to zero, which makes the volumetric
 equal to zero and leaving the sum of the surface forces equal zero. So, as δ →
forces
0, all4f aces Fi = 0 for i = 1, 2, 3 and ∴ τi A0 = τi1 A1 + τi2 A2 + τi3 A3 = τij Aj . But the area
 surface ⊥ to Xi is Ai = A0 ni . Therefore τi A0 = τij Aj = τij (A0 nj ), where τij Aj is
of each
the notation (represents the sum of all components). Thus τi = τij nj for i = 1, 2, 3,
where τi is the component of stress in the ith direction on a surface with a normal n . We
call τ i the stress vector and we call τij the stress matrix or tensor.

2
1.7.2 Example: Pascal’s Law for Hydrostatics
In a static fluid, the stress vector cannot be different for different directions of the surface
normal since there is no preferred direction in the fluid. Therefore, at any point in the
fluid, the stress vector must have the same direction as the normal vector n and the same
magnitude for all directions of n .

no summation
  
Pascal’s Law for hydrostatics: τij = − (pi ) (δij )
⎡ ⎤
−p1 0 0
τ = ⎣ 0 −p2 0 ⎦
0 0 −p3

where pi is the pressure acting perpendicular to the ith surface. If p0 is the pressure acting
perpendicular to the surface PQR, then τi = −ni p0 , but:

τi = τij nj = −(pi )δij nj = −(pi )(ni )


Therefore po = pi , i = 1, 2, 3 and n is arbitrary.

1.7.3 Symmetry of the Stress Tensor

To prove the symmetry of the stress tensor we follow the steps:

j
ij

ji ji

o i
ij

Figure 3: Material element under tangential stress.



1. The of surface forces = body forces + mass× acceleration. Assume no symmetry.
Balance of the forces in the ith direction gives:

(δ)(τij )T OP − (δ)(τij )BOT T OM = O(δ 2 ),

since surface forces are ∼ δ 2 , where the O(δ 2 ) terms include the body forces per unit
depth. Then, as δ → 0, (τij )T OP = (τij )BOT T OM .

2. The of surface torque = body moment + angular acceleration. Assume no sym­
metry. Balance of moments about o gives:

(τji δ)δ − (τij δ)δ = O(δ 3 ),

since the body moment is proportional to δ 3 . As δ → 0 , τij = τji .

1.8 Mass and Momentum Conservation


Consider a material volume ϑm and recall that a material volume is a fixed mass of mate­
rial. A material volume always encloses the same fluid particles despite a change in size,
position, volume or surface area over time.

1.8.1 Mass Conservation

The mass inside the material volume is:


M (ϑm ) = ρdϑ
ϑm(t)

Sm(t)

ϑm ( t )

Figure 4: Material volume ϑm (t) with surface Sm (t).

Therefore the time rate of increase of mass inside the material volume is:

d d
M (ϑm ) = ρdϑ = 0,
dt dt
ϑm (t)

which is the integral form of mass conservation for the material volume ϑm .

1.8.2 Momentum Conservation


The fluid velocity inside the material volume in the ith direction is denoted as ui . Linear
momentum of the material volume in the ith direction is

ρui dϑ
ϑm(t)

Newton’s law of motion: The time rate of change of momentum of the fluid in the material
control volume must equal the sum of all the forces acting on the fluid in that volume.
Thus:

d
(momentum)i =(body force)i + (surface force)i
dt
d
ρui dϑ = Fi dϑ + τij nj dS
dt 
ϑm (t) ϑm (t) Sm (t) τi


Divergence Theorems For vectors: ∇· vdϑ = ⊂⊃ 
v .n̂ dS
ϑ
∂vj
S vj nj
∂xj


∂τij
For tensors: dϑ = ⊂⊃ τij nj dS
∂xj
ϑ S

Using the divergence theorems we obtain


d ∂τij
ρui dϑ = Fi + dϑ
dt ∂xj
ϑm(t) ϑm(t)

which is the integral form of momentum conservation for the material volume ϑm .

1.8.3 Kinematic Transport Theorems

Consider a flow through some moving control volume ϑ(t) during a small time interval Δt.

Let f (x, t) be any (Eulerian) fluid property per unit volume of fluid (e.g. mass, momentum,

etc.). Consider the integral I(t):


I(t) = f (x, t) dϑ
ϑ(t)

According to the definition of the derivative, we can write

d I(t + Δt) − I(t)


I(t) = lim
dt Δt→0
⎧ Δt ⎫
⎪ ⎪
1 ⎨ ⎬
= lim f (x, t + Δt)dϑ − f (x, t)dϑ
Δt→0 Δt ⎪⎩ ⎪

ϑ(t+Δt) ϑ(t)

S(t+Δt)

ϑ ( t + Δt )

ϑ( t ) S(t)

Figure 5: Control volume ϑ and its bounding surface S at instants t and t + Δt.

Next, we consider the steps

1. Taylor series expansion of f (x, t + Δt) about (x, t).

∂f
f (x, t + Δt) = f (x, t) + Δt (x, t) + O((Δt)2 )
∂t

2. dϑ = dϑ + dϑ
ϑ(t+Δt) ϑ(t) Δϑ

where, dϑ = [Un (x, t)Δt] dS and Un (x, t) is the normal velocity of S(t).
Δϑ S(t)

S(t+Δt)

S(t)

v
U n ( x, t )Δt + O( Δt ) 2 dS

Figure 6: Element of the surface S at instants t and t + Δt.

Putting everything together:

⎧ ⎫

⎨ ⎪

d 1 ∂f 2
I(t) = lim dϑf + Δt dϑ + Δt dSUn f − dϑf + O(Δt) (1)
dt ⎪
Δt→0 Δt ⎩ ∂t ⎪

ϑ(t) ϑ(t) S(t) ϑ(t)

From Equation (1) we obtain the Kinematic Transport Theorem (KTT), which is equivalent
to Leibnitz rule in 3D.


d ∂f (x, t)
f (x, t)dϑ = dϑ + f (x, t)Un (x, t)dS
dt ∂t
ϑ(t) ϑ(t) S(t)

For the special case that the control volume is a material volume it is ϑ(t) = ϑm (t) and Un
= v · n
ˆ, where v is the fluid particle velocity. The Kinematic Transport Theorem (KTT),
then takes the form

d ∂f (x, t)
f (x, t)dϑ = dϑ + f (x, t)(v · n
ˆ)dS
dt ∂t   
ϑm (t) ϑm (t) Sm (t) f (vi ni )
(Einstein Notation)

Using the divergence theorem,



∇· α
dϑ = ⊂⊃ 
α
 ·n
ˆdS

ϑ

α S αi ni
∂xi i

we obtain the 1st Kinematic Transport Theorem (KTT)

 
d ∂f (x, t)
f (x, t) dϑ = + ∇ · (fv ) dϑ,
dt ∂t   
ϑm (t) ϑm (t) ∂
(f vi )
∂xi

where f is some fluid property per unit volume.

1.8.4 Continuity Equation for Incompressible Flow

• Differential form of conservation of mass for all fluids Let the fluid property
per unit volume that appears in the 1st KTT be mass per unit volume ( f = ρ):
 
d ∂ρ
0 = ρdϑ = + ∇ · (ρv ) dϑ
↑ dt ↑ ∂t
conservation ϑm (t) 1st KTT ϑm (t)
of mass
But since ϑm is arbitrary the integrand must be ≡ 0 everywhere.

Therefore:
∂ρ
+ ∇ · (ρv ) = 0
∂t
∂ρ
+ [v · ∇ρ + ρ∇ · v ] = 0
∂t  

Dt

Leading to the differential form of


Conservation of Mass: + ρ∇ · v = 0
Dt

10

• Continuity equation ≡ Conservation of mass for incompressible flow In


general it is ρ = ρ(p, T, . . .), but we consider the special case of an incompressible
flow, i.e. Dρ
Dt
= 0 (Lecture 2).

Note: For a flow to be incompressible, the density of the entire flow need not be
constant (ρ(x, t) =
const). As an example consider a flow of more than one incom­
pressible fluids, like water and oil, as illustrated in the picture below.

Constant ρ

fluid particle
ρ2

fluid particle oil water


ρ1
Figure 7: Interface of two fluids (oil-water)

Since for incompressible flows Dρ


Dt
= 0, substituting into the differential form of the
conservation of mass we obtain the

∂vi
Continuity Equation: ∇ · v ≡ = 0
∂xi
  
rate of volume dilatation

11

1.8.5 Euler’s Equation (Differential Form of Conservation of Momentum)

• 2nd Kinematic Transport Theorem ≡ 1st KTT + differential form of conservation


of mass for all fluids. If G = fluid property per unit mass, then ρG = fluid property
per unit volume

 
d ∂
ρGdϑ = (ρG) + ∇ · (ρGv ) dϑ
dt ↑ ∂t
ϑm (t) 1st KT T ϑm (t)
⎡ ⎤
⎢  ⎥
⎢ ∂ρ ∂G ⎥
⎢ G + ∇ · ρv +ρ + v · ∇G ⎥
= ⎢ ∂t ∂t ⎥ dϑ
⎣      ⎦
ϑm (t)
=0 from mass conservation = DG
Dt

The 2nd Kinematic Transport Theorem (KTT) follows:


d DG
ρGdϑ = ρ dϑ
dt Dt
ϑm ϑm

Note: The 2nd KTT is obtained from the 1st KTT (mathematical identity) and the
only assumption used is that mass is conserved.

12

• Euler’s Equation
We consider G as the ith momentum per unit mass (vi ). Then,

∂τij d Dvi
Fi + dϑ = ρvi dϑ = ρ dϑ
∂xj ↑ dt ↑ Dt
ϑm (t) conservation ϑm (t) 2nd KTT ϑm (t)
of momentum
But ϑm (t) is an arbitrary material volume, therefore the integral identity gives
Euler’s equation:

⎛ ⎞
Dvi ⎜ ∂vi ⎟ ∂τij
ρ ≡ ρ⎜
⎝ + v · ∇vi ⎟
⎠ = Fi +
Dt ∂t    ∂xj
∂v
vj ∂xi
j

And in vector tensor form:


Dv ∂v
ρ ≡ρ + v · ∇v = F + ∇ · τ
Dt ∂t

13

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 4

13.021 - Marine Hydrodynamics


Lecture 4

Introduction
Governing Equations so far:

Knowns Equations # Unknowns #

density ρ(x, t) Continuity 1 velocities vi (x, t) 3

(conservation of mass)


body force Fi Euler 3 stresses τij (x, t) 6

(conservation of momentum)

4 9


3 of the 9 unknowns of the stress tensor are eliminated by symmetry

The number of unknowns (9) is > than the number of equations (4), i.e. we don’t have
closure. We need constitutive laws to relate the kinematics vi to the dynamics τij .

1.9 Newtonian Fluids


1. Consider a fluid at rest (vi ≡ 0). Then according to Pascal’s Law:

τij = −ps δij (Pascal’s Law)


⎡ ⎤
−ps 0 0
τ = ⎣ 0 −ps 0 ⎦
0 0 −ps

where ps is the hydrostatic pressure and δij is the Kroenecker delta function, equal
to 1 if i = j and 0 if i =
 j.

2. Consider a fluid in motion. The fluid stress is defined as:

τij ≡ −pδij + τ̂ij


  
isotropic components all non-isotropic components
on diagonal both on/off diagonal
where p is the thermodynamic pressure and τ̂ij are the dynamic stresses. It should
be emphasized that −pδij includes all the isotropic components of the stress tensor
on the diagonal, while τ̂ij represents all the non-isotropic components, which may or
may not be on the diagonal (shear and normal stresses). The dynamic stresses τ̂ij is
related to the velocity gradients by empirical relations.
Experiments with a wide class of ‘Newtonian’ fluids showed that the dynamic stresses
are proportional to the rate of strain.

τ ij
Newtonian Fluid

Fluid

∂u k
∂x m

τ̂ij ≈ linear function of the ( rate of strain ≡ velocity gradient)




  
∂ ∂X ∂ ∂X ∂uk
=
∂t ∂x ∂x ∂t ∂xm
 
u

∂uk
i.e. τ̂ij ≈ αijkm i, j, k, m = 1, 2, 3
  ∂xm
34 =81
empirical coefficients
(constants for Newtonian fluids)

For isotropic fluids, this reduces to:





∂ui ∂uj ∂ul
τîj =μ + +λ ,
∂xj ∂xi ∂xl
 
∇·v

The fluid properties in the previous relation are:

• μ - (coefficient of) dynamic viscosity.


• λ - bulk elasticity, ‘second’ coefficient of viscosity

∂ui

For incompressible flow, we have shown that ∇ · v = = 0.


∂xi
Therefore, for an incompressible, isotropic, Newtonian fluid the dynamic or viscous
stresses τ̂ij are expressed as:


∂ui ∂uj
τ̂ij = μ +
∂xj ∂xi

1.9.1 Discussion on viscous stresses τ̂ij


1. Verify that for v = 0 we recover Pascal’s law.

∂ui

Proof: v = 0 ⇒ = 0 ⇒ τ̂ij = 0 ⇒ τij = −pδij + 0 ⇒ hydrostatic conditions


∂xj

2. Verify that τ̂ij = τ̂ji .




∂ui ∂uj
Proof: τ̂ij = μ + = τ̂ji ⇒ symmetry of stress tensor
∂xj ∂xi

3. When τ̂ij , i = j the viscous stress is a normal stress and is given by:

∂ui
τ̂ii = 2μ
∂xi

The normal viscous stresses τ̂ii are the diagonal terms of the viscous stress tensor.
The τ̂ii in general are not isotropic.

4. When τˆij , i =
 j the viscous stress is a shear stress and is given by:


∂ui ∂uj
τ̂ij = μ +
∂xj ∂xi

The shear viscous stresses τˆij , i = j are the off diagonal terms of the viscous stress
tensor.

⎡ ⎤
∂u ∂u ∂v
⎢ 2 ∂x +
∂y ∂x ⎥
5. A 2D viscous tensor has the form: μ


∂u ∂v

∂v
+ 2
∂y ∂x ∂y

6. Notation 1: The viscous stresses τ̂ij are often referred to (somewhat confusingly) as
shear stresses, despite the fact that when i = j the viscous stress is a normal stress.


∂ui ∂uj
7. Notation 2: Often τij ↔ τ̂ij and τij is used to denote μ ∂xj
+ ∂xi
.

4
1.10 Navier-Stokes equations(for Incompressible, Newtonian Fluid)

Equations # Unknowns #
Continuity 1 velocities vi (x, t) 3
Euler 3 stresses τ̂ij (x, t) 6
Newtonian 6 pressure p(x, t) 1
fluid symmetry
10 10
 
closure
To form the Navier-Stokes equations for incompressible, Newtonian fluids, we first substi­
tute the equation for the stress tensor for a Newtonian fluid, i.e.


∂ui ∂uj
τij = −pδij + τ̂ij = −pδij + μ +
↑ ∂xj ∂xi
Newtonian Fluid
into Euler’s equation:

Dui ∂τij
ρ = Fi +
Dt ∂xj


∂p ∂ ∂ui ∂uj
= Fi − +μ +
∂xi ∂xj ∂xj ∂xi

2


∂p ∂ ui ∂ ∂uj
= Fi − +μ +
∂xi ∂x2j ∂xi ∂xj
 
Continuity = 0

Therefore the Navier-Stokes equations for an incompressible, Newtonian fluid in cartesian


coordinates are given as:

Dui ∂ui ∂ui 1 ∂p ∂ 2 ui 1


= + uj =− + ν 2 + Fi Tensor form
Dt ∂t ∂xj ρ ∂xi ∂xj ρ

Dv ∂v 1 1

= + v · ∇v = − ∇p + ν∇2v + F Vector form


Dt ∂t ρ ρ
μ
where ν ≡ ρ
denoted as the kinematic viscosity [ L2 /T ].

5
• Unknowns and governing equations for incompressible, Newtonian fluids

Equations # Unknowns #

Continuity 1 pressure p(x, t) 1

Navier-Stokes 3 velocities vi (x, t) 3

4 4

• Values of constants (density, dynamic and kinematic viscosity) used in 13.021

water air units

density ρ 103 1 [kg/m3 ]

dynamic viscosity μ 10−3 10−2 [kg/ms]

kinematic viscosity ν 10−6 10−5 [m2 /s]

• Notation 1: The Continuity and the Navier-Stokes equations form the Governing
Equations for incompressible, Newtonian fluids.

Continuity + Navier-Stokes = Governing Equations

∂p
Notation 2: Alternatively, we refer to each equation Dv
Dt
i
= − ρ1 ∂xi
+ ν∇2 vi + ρ1 Fi
as the ith Momentum Equation. In this case, the Continuity and the Momentum
equations form the Navier-Stokes System of Equations.

Continuity + Momentum Equations = Navier-Stokes System of Equations

Both notations are equivalent, and in this text it will be made clear from the context
when the term Navier-Stokes refers to the Momentum Equations or to the System of
Governing Equations.

6
1.11 Boundary Conditions
In the previous paragraphs we formulated the governing equations that describe the flow
of an incompressible, Newtonian fluid. The governing equations (N-S) are a system of
partial differential equations (PDE’s). This 4 × 4 system of equations describes all the
incompressible flows, from rain droplets to surface waves.

One of the reasons this system of equations provides such different solutions lies on the
variety of the imposed boundary conditions. To complete the description of this problem
it is imperative that we specify appropriate boundary conditions. For the N-S equations
we need to specify ‘Kinematic Boundary Conditions’ and ‘Dynamic Boundary Conditions’.

1. Kinematic Boundary Conditions specify the boundary kinematics (position, ve­


locity, . . . ). On an impermeable solid boundary, velocity of the fluid = velocity of the
body. i.e. velocity continuity.

v = u ‘no-slip’ + ‘no-flux’ boundary condition


where v is the fluid velocity at the body and u is the body surface velocity

• v · n
ˆ = u · n ˆ ‘no flux’ ← continuous flow
• v · tˆ = u · t̂ ‘no slip’ ← finite shear stress

v
v

v
u

2. Dynamic Boundary Conditions specify the boundary dynamics ( pressure, sheer


stress, . . . ).

Stress continuity: p = p + p interface

τij = τij + τij , interface

The most common example of interfacial stress is surface tension.

p interface, τ ij interface
τij '
p'
τ ij
p

1.12 Surface Tension


• Notation: Σ [Tension force / Length] ≡ [Surface energy / Area].

• Surface tension is due to the intermolecular attraction forces in the fluid.

• At the interface of two fluids, surface tension implies in a pressure jump across the
interface. Σ gives rise to Δp across an interface.

• For a water/air interface: Σ = 0.07 N/m. This is a function of temperature, impurities


etc. . .

• 2D Example:
dθ dθ
cos · Δp · Rdθ = 2Σsin ≈ 2Σ dθ
2
  2  2

≈1 ≈ dθ
2
Σ
∴ Δp = R
Higher curvature implies in higher pressure jump at the interface.

8
p

p’=p+Δp dθ/2

Σ
R Σ

• 3D Example: Compound curvature




1 1
Δp = + Σ
R1 R2

where R1 and R2 are the principle radii of curvature.

1.13 Body Forces – Gravity


• Conservative forces

F = −∇ϕ for some ϕ,

where ϕ is the force potential.

  2  2
F · dx = 0 or F · dx = − ∇ϕ · dx = ϕ(x1 ) − ϕ(x2 )
1 1

• A special case of a conservative force is gravity F = −ρgk̂.

– In this case the gravitational potential is given by ϕg = ρgz. Therefore:

F = ∇(−ϕg ) = ∇(−ρgz) = ∇ps


 
≡ hydrostatic pressure ps

– Substitute in Navier-Stokes equation

Dv
ρ = −∇p + F
 +ρν∇2v
Dt
body force

= −∇p + ∇(−ρgz) + ρν∇2v

– Define: total pressure ≡ hydrostatic pressure + hydrodynamic pressure


p = ps + pd
p = −ρgz + pd =⇒

pd = p + ρgz

– Re-write Navier-Stokes:

Dv 
ρ = −∇ p + ρgz + ρν∇2v
Dt  
p d = p + ρgz
Dv
ρ = −∇pd + ρν∇2v
Dt
Therefore:
– Presence of gravity body force is equivalent to replacing the total pressure by a
dynamic pressure (pd = p − ps = p + ρgz) in the Navier-Stokes(N-S) equation.
– Solve the N-S equation with pd . To calculate the total pressure p simply add
back the hydrostatic component p = pd + ps = pd − ρgz.

10

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 5

13.021 - Marine Hydrodynamics


Lecture 5

Chapter 2 - Similitude (Keyword: EQUAL RATIOS)

Similitude: Similarity of behavior for different systems with equal similarity parameters.

Prototype ↔ Model

(real world) (physical/ analytical/ numerical . . . experiments)

Similitude Similarity Parameters (SP’s)


Geometric Similitude Length ratios, angles
Kinematic Similitude Displacement ratios, velocity ratios
Dynamic Similitude Force ratios, stress ratios, pressure ratios
..
.

Internal Constitution Similitude ρ, ν

Boundary Condition Similitude

..
.
For similitude we require that the similarity parameters SP’s (eg. angles, length ratios,
velocity ratios, etc) are equal for the model and the real world.

Example 1 Two similar triangles have equal angles or equal length ratios. In this case
the two triangles have geometric similitude.
Example 2 For the flow around a model ship to be similar to the flow around the prototype
ship, both model and prototype need to have equal angles and equal length
and force ratios. In this case the model and the prototype have geometric
and dynamic similitude.

1
2.1 Dimensional Analysis (DA) to Obtain SP’s
2.1.1 Buckingham’s π theory

Reduce number of variables → derive dimensionally homogeneous relationships.

1. Specify (all) the (say N) relevant variables (dependent or independent): x1 , x2 , . . . xN


e.g. time, force, fluid density, distance. . .

We want to relate the xi ’s to each other I( x1 , x2 , . . . xN ) = 0

2. Identify (all) the (say P) relevant basic physical units (“dimensions”)


e.g. M,L,T (P = 3) [temperature, charge, . . . ].
3. Let π = xα1 1 xα2 2 . . . xαNN be a dimensionless quantity formed from the xi ’s. Suppose

xi = Ci M mi Lli T ti , i = 1, 2, . . . , N
where the Ci are dimensionless constants. For example, if x1 = KE = 12 M V 2 =
1
2
M 1 L2 T −2 (kinetic energy), we have that C1 = 12 , m1 = 1, l1 = 2, t1 = −2. Then

π = (C1α1 C2α2 . . . CNαN )M α1 m1 +α2 m2 +...+αN mN Lα1 l1 +α2 l2 +...+αN lN T α1 t1 +α2 t2 +...+αN tN
For π to be dimensionless, we require
⎧ ⎫

⎪ 
 ⎪
N


⎪ ⎪



αi mi = 0 ⎪

P αi li = 0 aP × N system of Linear Equations (1)



⎪ ⎪


⎪ αi ti = 0 ⎪




 ⎪ ⎭

Σ notation
Since (1) is homogeneous, it always has a trivial solution,

αi ≡ 0, i = 1, 2, . . . , N (i.e. π is constant)
There are 2 possibilities:

(a) (1) has no nontrivial solution (only solution is π = constant, i.e. independent of
xi ’s), which implies that the N variable xi , i = 1, 2, . . . , N are Dimensionally Independent
(DI), i.e. they are ‘unrelated’ and ‘irrelevant’ to the problem.
(b) (1) has J(J > 0) nontrivial solutions, π1 , π2, . . . , πJ . In general, J < N , in fact,

J = N − K where K is the rank or ‘dimension’ of the system of equations (1).

2
2.1.2 Model Law
Instead of relating the N xi ’s by I(x1 , x2 , . . . xN ) = 0, relate the J π’s by

F (π1 , π2 , . . . πJ ) = 0, where J = N − K < N


For similitude, we require

(πmodel )j = (πprototype )j where j = 1, 2, . . . , J.


If 2 problems have all the same πj ’s, they have similitude (in the πj senses), so π’s serve
as similarity parameters.

Note:

• If π is dimensionless, so is π × const, π const , 1/π , etc. . .

• If π1 , π2 are dimensionless, so is π1 × π2 , π1
π2
, π1const1 × π2const2 , etc. . .

In general, we want the set (not unique) of independent πj ’s, for e.g., π 1 , π 2 , π 3 or π 1 , π 1
× π 2 , π 3 , but not π 1 , π 2 , π 1 × π 2 .

Example: Force on a smooth circular cylinder in steady, incompressible flow


Application of Buckingham’s π Theory.

ρ,ν
D

Figure 1: Force on a smooth circular cylinder in steady incompressible fluid (no gravity)

A Fluid Mechanician found that the relevant dimensional quantities required to evaluate
the force F on the cylinder from the fluid are: the diameter of the cylinder D, the fluid
velocity U , the fluid density ρ and the kinematic viscosity of the fluid ν. Evaluate the
non-dimensional independent parameters that describe this problem.

3
xi : F, U, D, ρ, ν → N = 5
xi = ci M mi Lli T ti → P = 3

N =5
 



F U D ρ ν
P = 3 mi 1 0 0 1 0
li 1 1 1 -3 2
ti -2 -1 0 0 -1

π = F α1 U α2 Dα3 ρα4 ν α5
For π to be non-dimensional, the set of equations

αi mi =0
αi li =0
αi ti =0

has to be satisfied. The system of equations above after we substitute the values for the
mi ’s, li ’s and ti ’s assume the form:

⎛ ⎞ α 1 ⎛ ⎞
1 0 0 1 0
⎜ α2 ⎟ 0

⎜ ⎟

1 1 1 −3 2

⎜ ⎟ ⎝


α3 ⎟
=
0

−2 −1 0 0 −1

α4 ⎠
0

α5
The rank of this system is K = 3, so we have j = 2 nontrivial solutions. Two families of
solutions for αi for each fixed pair of (α4 , α5 ), exists a unique solution for (α1 , α2 , α3 ).
We consider the pairs (α4 = 1, α5 = 0) and (α4 = 0, α5 = 1), all other cases are linear
combinations of these two.

1. Pair α4 = 1 and α5 = 0.
⎛ ⎞⎛ ⎞ ⎛ ⎞
1 0 0 α1 −1

⎝ 0 1 0 ⎠ ⎝ α2 ⎠ = ⎝ 4 ⎠

0 0 1 α3 2

which has solution


⎛ ⎞ ⎛ ⎞

α1 −1

⎝ α2 ⎠ = ⎝ 2 ⎠

α3 2

ρU 2 D2
∴ π1 = F α1 U α2 Dα3 ρα4 ν α5 =
F
Conventionally, π1 → 2π1−1 and ∴ π1 = 1
F
ρU 2 D2
≡ Cd , which is the Drag coefficient.
2

2. Pair α4 = 0 and α5 = 1.
⎛ ⎞⎛ ⎞ ⎛ ⎞
1 0 0 α1 0

⎝ 0 1 0 ⎠ ⎝ α2 ⎠ = ⎝ −2 ⎠

0 0 1 α3 −1

which has solution


⎛ ⎞ ⎛ ⎞

α1 0

⎝ α2 ⎠ = ⎝ −1 ⎠

α3 −1
ν
∴ π2 = F α1 U α2 Dα3 ρα4 υ α5 =
UD
Conventionally, π2 → π2−1 , ∴ π2 = UD
ν
≡ Re , which is the Reynolds number.
Therefore, we can write the following equivalent expressions for the non-dimensional inde­
pendent parameters that describe this problem:

F (π1 , π2 ) = 0 or π1 = f (π2 )
F (Cd , Re ) = 0 or Cd = f (Re )
F UD F UD
F( , )=0 or = f( )
1/2 ρU D
2 2 ν 1/2 ρU D
2 2 ν

5
Appendix A
Dimensions of some fluid properties

Quantities Dimensions
(M LT )
Angle θ none (M 0 L0 T 0 )
Length L L
Area A L2
Volume ∀ L3
Time t T
Velocity V LT −1
Acceleration V̇ LT −2
Angular velocity ω T −1
Density ρ M L−3
Momentum L M LT −1
Volume flow rate Q L3 T −1
Mass flow rate Q M T −1
Pressure p M L−1 T −2
Stress τ M L−1 T −2
Surface tension Σ M T −2
Force F M LT −2
Moment M M L2 T −2
Energy E M L2 T −2
Power P M L2 T −3
Dynamic viscosity μ M L−1 T −1
Kinematic viscosity ν L2 T −1

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 6

13.021 - Marine Hydrodynamics


Lecture 6

2.2 Similarity Parameters from Governing Equations and


Boundary Conditions
In this paragraph we will see how we can specify the SP’s for a problem that is governed by
the Navier-Stokes equations. The SP’s are obtained by scaling, non-dimensionalizing
and normalizing the governing equations and boundary conditions.
1. Scaling First step is to identify the characteristic scales of the problem.
For example: Assume a flow where the velocity magnitude at any point in space or
time |v(x, t)| is about equal to a velocity U , i.e. |v(
 x, t)| = α U , where α is such that
0 ≤ α ∼ O(1). Then U can be chosen to be the characteristic velocity of the flow
and any velocity v can be written as:

v = Uv 
where it is evident that v  is:
(a) dimensionless (no units), and
(b) normalized (|v  | ∼ O(1)).
Similarly we can specify characteristic length, time, pressure etc scales:
Characteristic scale Dimensionless and Dimensional quantity
normalized quantity in terms of characteristic scale
Velocity U v  v = Uv 
Length L x x = Lx
Time T t t = T t
Pressure po − pv p p = (po − pv )p

2. Non-dimensionalizing and normalizing the governing equations and bound­


ary conditions
Substitute the dimensional quantities with their non-dimensional expressions (eg.
substitute v with Uv  , x with Lx , etc) into the governing equations, and boundary
conditions. The linearly independent, non-dimensional ratios between the character­
istic quantities (eg. U , L, T , po − pv ) are the SP’s.

(a) Substitute into the Continuity equation (incompressible flow)

∇ · v = 0 ⇒
U  
∇ · v = 0 ⇒
L
∇ · v  = 0

Where all the () quantities are dimensionless and normalized (i.e., O(1)),
∂v 
for example, ∂x  = O(1).

(b) Substitute into the Navier-Stokes (momentum) equations

∂v 1
+ (v · ∇) v = − ∇p + ν∇2v − g ĵ ⇒
∂t ρ
U ∂v  U 2  po − pv   νU  2 
+ (v · ∇ ) v  = − ∇ p + 2 (∇ ) v − g ĵ
T ∂t  L ρU 2 L
U2
divide through by L , i.e., order of magnitude of the convective inertia term ⇒

� � � � � �
L ∂v p�
o − pv  �ν � 2 � gL
+ (v · ∇)v = − (∇p) + ∇ 
v − ĵ
U T ∂t ρU 2 UL U2

The coefficients ( � ) are SP’s.

2
Since all the dimensionless and normalized terms () are of O(1), the SP’s
( � ) measure the relative importance of each term compared to the con­
vective inertia. Namely,
∂v
L Eulerian inertia ∂t
• ≡ S = Strouhal number ∼ ∼
UT convective inertia (v · ∇)v

The Strouhal number S is a measure of transient behavior.

For example assume a ship of length L that has been travelling with velocity
U for time T . If the T is much larger than the time required to travel a ship
length, then we can assume that the ship has reached a steady-state.

L
<< T ⇒
U
L
= S << 1 ⇒
UT
∂v
ignore → assume steady-state
∂t

po − p v
• 1 ≡ σ = cavitation number.
2
ρU 2
The cavitation number σ is a measure of the likelihood of cavitation.

If σ >> 1, no cavitation. If cavitation is not a concern we can choose po


as a characteristic pressure scale, and non-dimensionalize the pressure p as
p = p o p
po pressure force
• 1 ≡ Eu = Euler number ∼
2
ρU 2 inertia force

UL inertia force
• ≡ Re = Reynold’s number ∼
ν viscous force
If Re >> 1, ignore viscosity.

U2 U � � 12
• = √ ≡ Fr = Froude number ∼ inertia force
gL gL gravity force

3
(c) Substitute into the kinematic boundary conditions

 boundary ⇒
u = U
u = U boundary


(d) Substitute into the dynamic boundary conditions

� ��
1 1 R=LR
p = pa + Δp = pa + + =⇒
R R p=(po −pv )p
� 1 �� 2 �
Δp
� � � �� �
 1 1 2 � /ρ
p = pa + 
+  = pa +
(po − pv ) L R1 R2 σ U 2L

U 2L inertial forces
• � ≡ We = Weber number ∼
/ρ surface tension forces

• Some SP’s used in hydrodynamics (the table is not exhaustive):

SP Definition

Reynold’s number Re UL
∼ inertia
ν viscous

Froude number Fr U2
∼ inertia
gL gravity
po pressure
Euler number Eu 1
ρU 2
∼ inertia
2

po −pv pressure
Cavitation number σ 1
ρU 2
∼ inertia
2

Strouhal number S L
∼ Eulerian inertia
UT convective inertia
Weber number We U 2L
∼ inertia
Σ/ρ surface tension

2.3 Similarity Parameters from Physical Arguments


Alternatively, we can obtain the same SP’s by taking the dimensionless ratios of significant
flow quantities. Physical arguments are used to identify the significant flow quantities.
Here we obtain SP’s from force ratios. We first identify the types of dominant forces acting
on the fluid particles. The SP’s are merely the ratios of those forces.

1. Identify the type of forces that act on a fluid particle:


� 2�
1.1 Inertial forces ∼ mass × acceleration ∼ (ρL ) UL = ρU 2 L2
3

∂u � U� 2
1.2 Viscous forces ∼ μ × area ∼ μ L (L ) = μU L
∂y
����
shear stress

1.3 Gravitational forces ∼ mass × gravity ∼ (ρL3 )g

1.4 Pressure forces ∼ (po − pv )L2

2. For similar streamlines, particles must be acted on forces whose resultants are in the
same direction at geosimilar points. Therefore, the following force ratios must be
equal:

inertia ρU 2 L2 UL
• ∼ = ≡ Re
viscous μU L ν

� �1/2 � �1/2
inertia ρU 2 L2 U
• ∼ =√ ≡ Fr
gravity ρgL3 gL

�1 �−1
2
inertia (po − pv )L2 p o − pv
• ∼ 1 2 2
= 1 2 ≡σ
pressure 2
ρU L 2
ρU

2.4 Importance of SP’s


• The SP’s indicate whether different systems have similar flow properties.

• The SP’s provide guidance in approximating complex physical problems.

Example A hydrofoil of length L is submerged in a known fluid (density ρ, kinematic


viscosity ν). Given that the hydrofoil is travelling with velocity U and the gravitational
acceleration is g, determine the hydrodynamic force F on the hydrofoil.

F ρ,ν
g U
L

SP’s for this problem:

L p o − pv U 2L U UL
S= , σ= 1 , We = � , Fr = √ , Re =
UT 2
ρU 2 /ρ gL ν

We define the dimensionless force coefficient:


F
CF ≡ 1
2
ρU 2 L2
The force coefficient must depend on the other SP’s:
� � �
CF = CF (S, σ, We , Fr , Re ) or CF = CF S, σ −1 , We−1 , Fr , Re−1

Procedure We will first study under what conditions each SP → 0.


We will estimate CF for the case that all of the SP’s → 0.

1. Significance of the Strouhal number S = L/U T .

Change S keeping all other SP’s (σ −1 , We−1 , Fr , Re−1 ) fixed.

CF transient

Steady-State

S-1 = UT / L
S~O(1)
S~O( 1)

Exact position of the cut


depends on the problem and
the quantities of interest.


For S << 1, assume steady-state: ∂t =0
For S >> 1, unsteady effect is dominant.

For example, for the case L = 10m and U = 10m/s we can neglect the unsteady
effects when:

L L
S << 1 ⇒ << 1 ⇒ T >> ⇒ T >> 1s
UT U

Therefore for T >> 1s we can approximate S 1 and we can assume steady state.
In the case of a steady flow:

� �
CF = CF S 0, σ −1 , We−1 , Fr , Re−1 ⇒

� �
CF ∼
= CF σ −1 , We−1 , Fr , Re−1

7
po − pv
2. Significance of the cavitation number σ = 1 .
2
ρU 2

Change σ −1 keeping all other SP’s (S 0, We−1 , Fr , Re−1 ) fixed.

Some comments on cavitation


pv : Vapor pressure, the pressure at which water boils
p o ≤ pv : State of fluid changes from liquid to gas ⇒ CAVITATION
Consequences : Unsteady → Vibration of structures, which may lead to fatigue, etc
Unstable → Sudden cavity collapses
→ Large force acting on the structure surface
→ Surface erosion

CF

Strong
cavitation No cavitation

σinception σ

For σ << 1, cavitation occurs.

For σ >> 1 ⇒ σ −1 << 1, cavitation will not occur.

In general cavitation occurs when we have large velocities, or when po ∼ pv

8
For example, assume a hydrofoil travelling in water of density ρ = 103 kg/m3 .

The characteristic pressure is po = 105 N/m2 and the vapor pressure is pv = 103 N/m2 .
Cavitation will not occur when:

1 2
2
ρU p o − pv
σ −1 << 1 ⇒ << 1 ⇒ U << 1 ⇒ U << 14m/s
p o − pv 2
ρ

Therefore for U << 14m/s it is σ >> 1 ⇒ σ −1 0 and cavitation will not occur.

In the case of a steady, non-cavitating flow:

� �
CF = CF 0, σ −1 0, We−1 , Fr , Re−1 ⇒

� �
CF ∼
= CF We−1 , Fr , Re−1

U 2L
3. Significance of the Weber number We = .
Σ/ρ

Change We−1 keeping the other SP’s (S 0, σ −1 0, Fr , Re−1 ) fixed.

For We << 1, surface tension is significant.


For We >> 1 ⇒ We−1 << 1, surface tension is not significant.

For example, assume a hydrofoil travelling with velocity U = 1m/s� near an air/water
interface (water density ρ = 103 kg/m3 , surface tension coefficient = 0.07N/m).

Surface tension can be neglected when:


Σ Σ
ρ ρ
We−1 << 1 ⇒ << 1 ⇒ L >> ⇒ L >> 7 · 10−5 m
U 2L U2

Therefore for L >> 7 · 10−5 m it is We >> 1We−1 1 and surface tension effects can
be neglected.

So in the case of a steady, non-cavitating, non-surface tension flow:

� �
CF = CF 0, 0, We−1 0, Fr , Re−1 ⇒

� �
CF ∼
= CF Fr , Re−1

10

U
4. Significance of the Froude number Fr = √ , which measures the ‘gravity effects’.
gh

Change Fr keeping the other SP’s (S 0, σ −1 0, We−1 0, Re−1 ) fixed.

‘Gravity effects’, hydrostatic pressure do not create any flow (isotropic) nor do they
change the flow dynamics unless Dynamic Boundary Conditions apply.

√ √
‘Gravity effects’ are not significant when U << gh ⇒ Fr 0, or U >> gh ⇒
Fr−1 0. Physically, this is the case when the free surface is

• absent or
• far away or
• not disturbed, i.e., no wave generation.

The following figures (i - iv) illustrate cases where gravity effects are not significant.

11

In any of those cases the gravity effects are insignificant and equivalently Fr is not
important (i.e. Fr 0 or Fr−1 0).

So in the case of a steady, non-cavitating, non-surface tension, with no gravity


effects flow:

� �
CF = CF 0, 0, 0, Fr 0 or Fr−1 0, Re−1 ⇒

� �
CF ∼
= CF Re−1

A look ahead: Froude’s Hypothesis

Froude’s Hypothesis states that

CF = CF (Fr , Re ) = C1 (Fr ) + C2 (Re )

Therefore dynamic similarity requires

(Re )1 = (Re )2 , and


(Fr )1 = (Fr )2

Example: Show that if ν and g are kept constant, two systems


(1, 2) can be both geometrically and dynamically similar only
if:

L1 = L2 , and
U1 = U 2

12

UL
5. Significance of the Reynolds number Re = .
ν

Change Re keeping the other SP’s (S 0, σ −1 0, We−1 0, Fr 0 or Fr−1 0)


fixed.

Recall that for a steady, non-cavitating, non-surface tension, with no gravity


effects flow: � �
CF = CF Re−1

CF Sphere

Plate

Laminar (Re)cr Turbulent Re


Transition
Re << 1, Stokes flow (creeping flow)
Re < (Re )cr , Laminar flow
Re > (Re )cr , Turbulent flow
Re → ∞, Ideal fluid

For example, a hydrofoil of cord length L = 1m travelling in water (kinematic vis­


cosity ν = 10−1 m2 /s) with velocity U = 10m/s has a Reynolds number with respect
to L:

ν
Re = = 107 → ideal fluid, and Re−1 0
UL

13
Therefore for a steady, non-cavitating, non-surface tension, with no-gravity
effects flow in an ideal fluid:

CF = CF (0, 0, 0, 0, 0) = constant = 0

→D’Alembert’s Paradox
No drag force on moving body in ideal fluid.

14

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 7

13.021 - Marine Hydrodynamics


Lecture 7

Chapter 3 – Ideal Fluid Flow

The structure of Lecture 7 has as follows: In paragraph 3.0 we introduce the concept of
inviscid fluid and formulate the governing equations and boundary conditions for an ideal
fluid flow. In paragraph 3.1 we introduce the concept of circulation and state Kelvin’s
theorem (a conservation law for angular momentum). In paragraph 3.2 we introduce the
concept of vorticity.


⎪ Inviscid Fluid ν=0

Ideal Fluid Flow ≡ +

⎩ Incompressible Flow (§ 1.1) Dρ
= 0 or ∇ · v = 0
Dt

3.0 Governing Equations and Boundary Conditions for Ideal Flow


• Inviscid Fluid, Ideal Flow
Recall Reynolds number is a qualitative measure of the importance of viscous forces
compared to inertia forces,

UL inertia forces
Re = =
ν viscous forces
For many marine hydrodynamics problems studied in 13.021 the characteristic lengths
and velocities are L ≥ 1m and U ≥ 1m/s respectively. The kinematic viscosity in
water is νwater = 10−6 m2 /s leading thus to typical Reynolds numbers with respect to
U and L in the order of

UL
Re = ≥ 106 >>> 1 ⇒
ν
1 viscous forces
∼ 0
Re inertia forces

This means that viscous effects are << compared to inertial effects - or confined
within very small regions. In other words, for many marine hydrodynamics prob­
lems, viscous effects can be neglected for the bulk of the flow.

Neglecting viscous effects is equivalent to setting the kinematic viscosity ν = 0, but

ν = 0 ⇔ inviscid fluid

Therefore, for the typical marine hydrodynamics problems we assume

incompressible flow + inviscid fluid ≡ ideal fluid flow

which turns out to be a good approximation for many problems.

• Governing Equations for Ideal Fluid Flow

– Continuity Equation:

∇ · v = 0

– Momentum (Navier-Stokes ⇒ Euler) equations:

∂v 1
+ v · ∇v = − ∇p − g ĵ
∂t ρ

By neglecting the viscous stress term (ν∇2v ) the Navier-Stokes equations reduce

to the Euler equations. (Careful not to confuse this with the Euler equation in

§1.6).

The N-S equations are second order PDE’s with respect to space

(2nd order in ∇2 ), thus: (a) require 2 kinematic boundary conditions, and (b)

produce smooth solutions in the velocity field.

The Euler equations are first order PDE’s, thus: (a) require 1 kinematic bound­

ary condition, and (b) may allow discontinuities in the velocity field.

• Boundary Conditions for Euler equations (Ideal Flow):

– KBC:
� �
v · n̂ = u · n
ˆ=U ← ‘no flux’ + free (to) slip
  n
given

 · t̂ does not apply.


Note: ‘No slip’ condition v · tˆ = U

The ‘no slip’ condition is required to ensure that the velocity gradients are finite
and therefore the viscous stresses τ̂ij are finite.

But since ν = 0 the viscous stresses are identically zero (τ̂ij ∝ μ = ρν = 0)


and the velocity gradients can be infinite. Or else the velocity field need not be
continuous.

∂u
Viscousflow τ w ∝ μ
y u(dy) ∂y
<∞ <∞

u(0) ∂u
U Inviscidflow τ w ∝0
∂y

– DBC:

p = . . . Pressure given on the boundary

Similarly to the argument for the KBC, viscous stresses τ̂ij cannot be specified
on any boundary since ν = 0.

• Summary of consequences neglecting viscous effects this far:

– Neglecting viscous effects is equivalent to setting the kinematic viscosity equal


to zero:

ν=0

Setting ν = 0 ⇐⇒ inviscid fluid

– Setting ν = 0 the viscous term in the Navier-Stokes equations drops out and we
obtain the Euler equations.

The Euler equations are 1st order PDE’s in space, thus (a) require only one
boundary condition for the velocity and (b) may allow for velocity jumps.

– Setting ν = 0 all the viscous stresses τ̂ij ∝ μ = ρν are identically 0. This may
allow for infinite velocity gradients.

This affects (a) the KBC, allowing free slip, and (b) the DBC, where no viscous
stresses can be specified on any boundary.

4
3.1 Circulation – Kelvin’s Theorem
3.1.1 Γ ≡ Instantaneous circulation around any arbitrary closed contour C.

C v
v v
dx

Γ= v 
· dx

C tangential
velocity

The circulation Γ is an Eulerian idea and is instantaneous, a ‘snapshot’.

3.1.2 Kelvin’s Theorem (KT) :


For ideal fluid under conservative body forces,


= 0 following any material contour C,
dt
i.e., Γ remains constant under for Ideal Fluid under Conservative Forces (IFCF).

This is a statement of conservation of angular momentum.

(Mathematical Proof: cf JNN pp 103)

Kinematics of a small deformable body:

(a) Uniform translation → Linear momentum


(b) Rigid body rotation → Angular momentum
(c) Pure strain→ No linear or angular momentum involved (no change in volume
(d) Volume dilatation

For Ideal Fluid under Conservative body Forces:

(a) Linear momentum → Can change


(b) Angular momentum → By K.T., cannot change
(c) Pure strain→ Can change
(d) Volume dilatation → Not allowed (incompressible fluid)

Kelvin’s Theorem is a statement of conservation of angular momentum


under IFCF.

Example 1: Angular momentum of point mass.

v1
v2
r1 m1
θ r2 m2

Angular momentum of point mass:





L
= |r × (mv )| = mvr = mr
2 θ̇

Conservation of angular momentum:





L


=

L

1 2
1 2 m =m
m1 v1 r1 = m2 v2 r2 =⇒
v1 r1 = v2 r2 or
r12 θ̇1 = r22 θ̇2

Conservation of angular momentum does not imply constant angular


velocity:

Angular Momentum � angular velocity ω




Example 2: Conservation of circulation around a shrinking circular material volume


Vm .

υ2
υ1 r1
r2

Vm
Vm




Γ1 = dθr1 v1 = dθr2 v2 = Γ2
0 0

Example 3: Conservation of circulation around a shrinking arbitrary material vol­


ume Vm , Cm .

Γ1 = v1 · dx = v2 · dx = Γ2


C1 C2

3.2 Vorticity
3.2.1 Definition of Vorticity
  
∂w ∂v ∂w ∂u ∂v ∂u
ω
 = ∇ × v = − î − − ĵ + − k̂
∂y ∂z ∂x ∂z ∂x ∂y

Relationship of vorticity to circulation - Apply Stokes’ Theorem:

Γ= v · dx = (∇ × v) · ndS


ˆ = ω · ndS
ˆ ≡ Flux of vorticity out of S
C any S ‘covering’ C any S ‘covering’ C

3.2.2 What is Vorticity?



For example, special case: 2D flow - w = 0; ∂z
= 0; ωy = ωx = 0 and

∂v ∂u
ωz = −
∂x ∂y

(a) Translation: u = constant, v = constant

time t + Δt

uiˆ + υ ˆj uiˆ + υ ˆj

uiˆ + υ ˆj uiˆ + υ ˆj
time t

∂v ∂u
= 0, = 0 ⇒ ωz = 0 → no vorticity
∂x ∂y

9
(b) Pure Strain (no volume change):

Areat + Δt

Areat

No volume change Areat = Areat + Δt

∂u ∂v ∂u ∂v
= − ; u = -v; = 0; = 0 ⇒ ωz = 0
∂x ∂y ∂y ∂x

10

(c) Angular deformation

δ x = ⎛⎜ ∂u ∂y dy ⎞⎟Δt
⎝ ⎠

r
υ = ∂u ∂y dy iˆ + ∂υ ∂y dy ˆj time
t + Δt

dy

time t
( )
δ y = ∂υ ∂x dx Δt
dx
r r
υ =0 υ = ∂u ∂x dx iˆ + ∂υ ∂x dx ĵ

∂u ∂v
 = 0 only if
ω = → δx = δy( for dx = dy)
∂y ∂x

11

(d) Pure rotation with angular velocity Ω

time t + Δt r
υ = − Ωdy iˆ
Ω Δt time t

dy

Ω Δt

Ω r
υ =0
dx r
υ = Ωdx ĵ

∂v ∂u
= Ω; = −Ω; ωz = 2Ω
∂x ∂y
i.e. vorticity ∝ 2(angular velocity).

3.2.3 Irrotational Flow


A flow is irrotational if the vorticity is zero everywhere or if the circulation is zero
along any arbitrary closed contour:


ω ≡ 0 everywhere ⇔ Γ ≡ 0 for any C
Further on, if at t = to , the flow is irrotational, i.e., Γ ≡ 0 for all C, then Kelvin’s
theorem states that under IFCF, Γ ≡ 0 for all C for all time t:

once irrotational, always irrotational

(Special case of Kelvin’s theorem)

12
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 8

13.021 - Marine Hydrodynamics


Lecture 8
In Lecture 8, paragraph 3.3 we discuss some properties of vortex structures. In paragraph
3.4 we deduce the Bernoulli equation for ideal, steady flow.

3.3 Properties of Vortex Structures


3.3.1 Vortex Structures

• A vortex line is a line everywhere tangent to ω


.
vortex line
r
ω2
Ω2 r
u2

r
ω1
Ω1 r
u1

• A vortex tube (filament) is a bundle of vortex lines.


vortex tube

r
u
r
ω

vortex lines

1
• A vortex ring is a closed vortex tube.
A sketch and two pictures of the production of vortex rings from orifices are

shown in Figures 1, 2, and 3 below.

(Figures 2,3: Van Dyke, An Album of Fluid Motion 1982 p.66, 71)

side view
v
u v
u
U v
ω
v
u Γ

cross section
v v
u
ω
U
v v
u ω

Figure 1: Sketch of vortex ring production

3.3.2 No Net Flux of Vorticity Through a Closed Surface


Calculus identity, for any vector v :

∇ · (∇ × v) = 0 ⇒
 
ω

∇ · ω =0⇒
 
∇ · ω =  ω
 · n̂
 dS = 0

V Divergence S vorticity flux

Theorem

i.e. The net vorticity flux through a closed surface is zero.

(a) No net vorticity flux through a vortex tube:

(Vorticity Flux)in = (Vorticity Flux)out ⇒ (ω · n̂)in δAin = (ω


 · n̂)out δAout

(ωv ⋅ n̂ )out

v
ω ⋅ nˆ = 0
(ωv ⋅ n̂ )in

(b) Vorticity cannot stop anywhere in the fluid. It either traverses the fluid begin­
ning or ending on a boundary or closes on itself (vortex ring).

r r
ω ω

3.3.3 Conservation of Vorticity Flux


 
0 = Γ3 = v · dx = ω
 · ndS
ˆ =0
C3 S3

n̂1 C3 n̂2
C1 C2

  
Γ1 = v · dx = ω
 · n̂1 dS = ω
 · n̂2 dS = Γ2
C1 S1 S2

Therefore, circulation is the same in all circuits embracing the same vortex tube. For
the special case of a vortex tube with ‘small’ area:

Γ = ω1 A1 = ω2 A2

ω1 ω2

A1 A2

An application of the equation above is displayed in the figure below:

ω2
ω1
A2 = 2A1
A1
ω2 = ω1/2

3.3.4 Vortex Structures are Material Structures


Consider a material patch Am on a vortex tube at time t.

Am

∂Am

By definition,

ω · n̂ = 0 on An

Then,
 
Γ∂Am = v · dx = ω
 · nds
ˆ =0
∂Am Am

At time t + Δt, Am moves, and for an ideal fluid under the influence of conservative
body forces, Kelvin’s theorem states that

Γ∂Am = 0
So, ω · n̂ = 0 on Am still, i.e., Am still on the vortex tube. Therefore, the vortex tube
is a material tube for an ideal fluid under the influence of conservative forces. In the
same manner it can be shown that a vortex line is a material line, i.e., it moves with
the fluid.

3.3.5 Vortex stretching


Consider a small vortex filament of length L and radius R, where by definition ω is
tangent to the tube.

R A

Γ = ωA = constant (in time)


↑ ↑
Stokes Kelvins’
Theorem Theorem

But tube is material with volume = AL = πR2 L = constant in time (continuity)

Γ ωA ω
∴ = = = constant
Volume LA L
As a vortex stretches, L increases, and since the volume is constant (from continuity),
A and R decrease, and due to the conservation of the angular momentum, ω increases.
In other words,

Vortex stretching ⇔ L ↑ ⇒ ω ↑ (conservation of angular momentum)


⇒ A and R ↑ (continuity)

3.3.6 Summary on Vortex Structures


r
ω
Γ Γ

R
r
r
ω A

Vortex ring length L = 2πR [L]

Cross sectional area A = πr2 [L2 ]

Vortex ring volume ∀ = AL = const [L3 ]



continuity

Vorticity ω = ∇ × v [T−1 ]

Circulation Γ = const [L2 T−1 ]



Kelvin’s theorem

Γ = ωA = const [L2 T−1 ]



vorticity flux through A

Γ ∝ U r = const [L2 T−1 ]

Continuity relates length ratios



⎪ ∀

⎪ A ∝ ∴ as L ↑ A↓

⎪ L


∀ = LA = const






⎪ ∀

⎩ r ∝ ∴ as L ↑ r↓
L

Kelvin’s theorem + Continuity relate length ratios to Γ, ω, U

Γ r∝ ∀/L L
U r ∝ Γ = const → U∝ → U ∝Γ ∴ as L ↑ U↑
r ∀

Γ A∝∀/L L
ωA ∝ Γ = const → ω∝ → ω∝Γ ∴ as L ↑ ω↑
A ∀

Example 1: Example 2:

A2 ω2
r A2

A1
Γ Γ2
r
L2 r
ω1 ω2
L1
A1 r Γ1
ω1

A1 < A2 L1 < L 2 Given


ω1 > ω 2 A1 > A2 From continuity only
Γ1 = Γ2 From Kelvin’s theorem
ω1 < ω 2 From Kelvin’s theorem + continuity
U1 < U 2 From Kelvin’s theorem + continuity

8

3.4 Bernoulli Equation for Steady ( ∂t = 0), Ideal(ν = 0),
Rotational flow


p=f v Viscous flow: Navier-Stokes’ Equations (Vector Equations)
p = f (|v |) Ideal flow: Bernoulli Equation (Scalar equation)

Steady, inviscid Euler equation (momentum equation):


 
p
v · ∇v = −∇ + gy (1)
ρ
From Vector Calculus we have

∇ (u · v ) = (u · ∇) v + (v · ∇) u + u × (∇ × v ) + v × (∇ × u) ⇒


∇( 12 |v |2 ) = v · ∇v + v × (∇ × v ) ⇒
 2
v
v · ∇v = ∇ − v × (∇ × v ) where v 2 ≡ v · v = |v |2
2

From the previous identity and Equation (1) we obtain


 2  
v p
v · (1) → v · ∇ − v · v × (∇ × v ) = −v · ∇ + gy
2    ρ
 ⊥v 
0
v · momentum (1) → energy

Therefore,    
v2 p D v2 p
v · ∇ 2
+ ρ
+ gy = 0 = Dt 2
+ ρ
+ gy
streamline pathline
v2 p
i.e., 2
+ ρ + gy = constant on a streamline
v2 p
In general, 2
+ ρ
+ gy = F (Ψ) where Ψ is a tag for a particular streamline.

Assumptions: Ideal fluid, Steady flow, Rotational in general.

3.4.1 Example: Contraction in Water or Wind Tunnel

Contraction Ratio: γ = R1 /R2 >> 1 ( γ = O(10) for wind tunnel ; γ = O(5) for water
tunnel)

Let U¯1 and U¯2 denote the average velocities at sections 1 and 2 respectively.
 2
2 2 U¯2 R1
1. From continuity: Ū1 πR1 = Ū2 πR2 → ¯ = = γ 2 >> 1
U1 R2
2.

∂u
Since 0,ω
=  = 0 → vortex ring.
∂r

10


⎪ ω1 ω2 ω2 R2 1

⎪ = → = ∼ << 1

2πR1 2πR2 ω1 R1 γ
ω/L = constant ⇒
   

⎪ ∂u ∂u ∂u

⎪ since ω ∼ → <<

∂r ∂r 2 ∂r 1

i.e.,

Section 2
Section 1

3. Near the center, let U1 = U¯1 (1 + ε1 ) and U2 = U¯2 (1 + ε2 ) where ε1 and ε2 measure
the relative velocity fluctuations. Apply the Bernoulli equation along a reference
average streamline

P1 + 12 ρŪ12 = P2 + 12 ρŪ22 (2)

Apply Bernoulli Equation to a particular streamline


 2 
2
P1 + 12 ρ U¯1 (1 + ε1 ) = P2 + 12 ρ
Ū2 (1 + ε2 )

From (2) and (3) we obtain

ε2 Ū 2 1
ε1 Ū12 = ε2 Ū22 + O(ε2 ) → ∼ 12 ∼ 4 << 1
ε1 Ū2 γ

11

13.021 – Marine Hydrodynamics, Fall 2004


Lecture 9

13.021 - Marine Hydrodynamics


Lecture 9
Lecture 9 is structured as follows: In paragraph 3.5 we return to the full Navier-Stokes
equations (unsteady, viscous momentum equations) to deduce the vorticity equation and
study some additional properties of vorticity. In paragraph 3.6 we introduce the concept of
potential flow and velocity potential. We formulate the governing equations and boundary
conditions for potential flow and finally introduce the stream function.

3.5 Vorticity Equation


Return to viscous incompressible flow. The Navier-Stokes equations in vector form
 
∂v p
+ v · ∇v = −∇ + gy + ν∇2v
∂t ρ
By taking the curl of the Navier-Stokes equations we obtain the vorticity equation. In
detail and taking into account ∇ × u ≡ ω  we have

 
∂v p  
∇ × (Navier-Stokes) → ∇ × + ∇ × (v · ∇v ) = −∇ × ∇ + gy + ∇ × ν∇2v
∂t ρ
The first term on the left side, for fixed reference frames, becomes
∂v ∂ ∂ω
∇× = (∇ × v ) =
∂t ∂t ∂t
In the same manner the last term on the right side becomes
 
∇ × ν∇2v = ν∇2 ω 
Applying the identity ∇ × ∇ · scalar = 0 the pressure term vanishes, provided that the
density is uniform  
p
∇ × ∇( + gy) = 0
ρ

1
The inertia term v · ∇v , as shown in Lecture 8, §3.4, can be rewritten as
 2
1 v
v · ∇v = ∇ (v · v ) − v × (∇ × v ) = ∇  where v 2 ≡ |v |2 = v · v
− v × ω
2 2

and then the second term on the left side can be rewritten as
 2
v
∇ × (v · ∇) v = ∇ × ∇ − ∇ × (v × ω ) = ∇ × (ω × v )
2
= (v · ∇) ω − (ω · ∇) v + ω (∇ · v ) + v (∇ · ω )
   
=0 =0 since
incompressible ∇·(∇×v )=0
fluid

Putting everything together, we obtain the vorticity equation

Dω
 · ∇) v + ν∇2 ω
= (ω 
Dt

Comments-results obtained from the vorticity equation

• Kelvin’s Theorem revisited - from vorticity equation:

D
ω
If ν ≡ 0, then Dt
= (ω · ∇) v , so if ω
 ≡ 0 everywhere at one time, ω
 ≡ 0 always.

• ν can be thought of as diffusivity of vorticity (and momentum), i.e., ω once generated


(on boundaries only) will spread/diffuse in space if ν is present.

v
v ω
ω
v v
Dv v Dω
D ω v
= υ∇ 2v + ... = υ∇ 2ω + ...
Dt Dt

∂T
• Diffusion of vorticity is analogous to the heat equation: = K∇2 T , where K is the
∂t
heat diffusivity.

2
Numerical example  ν ∼ 1 mm /s. For diffusion time t = 1 second, diffusion
√ for
distance L ∼ O νt ∼ O (mm). For diffusion distance L = 1cm, the necessary
diffusion time is t ∼ O (L2 /ν) ∼ O(10)sec.

• In 2D space (x, y),


v = (u, v, 0) and ≡0
∂z

So, ω
 = ∇ × v is ⊥ to v (ω
 is parallel to the z-axis). Then,

⎛ ⎞
⎜ ∂ ∂ ∂ ⎟
(ω · ∇) v = ⎝ ωx + ωy + ωz ⎠ v ≡ 0,
 ∂x  ∂y ∂z
0 0

0

so in 2D we have

Dω
= ν∇2 ω

Dt
If ν = 0, D
ω
Dt
= 0, i.e., in 2D following a particle the angular velocity is conserved.
Reason: In 2D space the length of a vortex tube cannot change due to continuity.

• In 3D space,

Dωi ∂vi ∂ 2 ωi
= ωj + ν
Dt ∂x ∂xj ∂xj
  j  
vortex turning and stretching diffusion

for example,

Dω2 ∂u2 ∂u2 ∂u2


= ω1 + ω2 + ω3 + diffusion
Dt ∂x1 ∂x ∂x3
    2  
vortex turning vortex stretching vortex turning

z ≡ x3 z ≡ x3
∂u2
dz > 0
dz ∂x3
dy
u2 = 0 y ≡ x2 u2 = 0 y ≡ x2
∂u2
dy > 0
∂x2
x ≡ x1 x ≡ x1

∂u2 Dω2 ∂u2 ω3 >0 Dω2


>0 ⇒ >0 > 0 ⎯⎯ ⎯→ >0
∂x2 1Dt
424 3 ∂x3 1Dt
424 3
vortex stretching rate vortex turning rate

3.5.1 Example: Pile on a River

Scouring

What really happens as length of the vortex tube L increases?

IFCF is no longer a valid assumption.

Why?
Ideal flow assumption implies that the inertia forces are much larger than the viscous
effects. The Reynolds number, with respect to the vortex tube diameter D is given by
UD
Re ∼
ν
As the vortex tube length increases ⇒ the diameter D becomes really small ⇒ Re is not
that big after all.
Therefore IFCF is no longer valid.

3.6 Potential Flow


Potential Flow (P-Flow) is an ideal and irrotational fluid flow

⎧ ⎫

⎪ Inviscid Fluid ν=0 ⎬


⎨ + Ideal Flow

P-Flow ≡ Incompressible Flow ∇ · v = 0



⎪ +

Irrotational Flow ω = 0 or Γ = 0

3.6.1 Velocity potential

For ideal flow under conservative body forces by Kelvin’s theorem if ω ≡ 0 at some
time t, then ω ≡ 0 ≡ irrotational flow always. In this case the flow is P-Flow.

Given a vector field v for which ω  = ∇ × v ≡ 0, there exists a potential function


(scalar) - the velocity potential - denoted as φ, for which

v = ∇φ

Note that
ω
 = ∇ × v = ∇ × ∇φ ≡ 0
for any φ, so irrotational flow guaranteed automatically. At a point x and time t,
the velocity vector v (x, t) in cartesian coordinates in terms of the potential function
φ(x, t) is given by
 
∂φ ∂φ ∂φ
v (x, t) = ∇φ (x, t) = , ,
∂x ∂y ∂z

φ (x)

u u

∂φ u=0 ∂φ
>0 <0
∂x ∂x
u >0 u <0

from low φ ⎯
⎯→ to high φ

The velocity vector v is the gradient of the potential function φ, so it always points
towards higher values of the potential function.

3.6.2 Governing Equations and Boundary Conditions for Potential Flow

(a) Continuity

∇ · v = 0 = ∇ · ∇φ ⇒ ∇2 φ = 0

Number of unknowns → φ

Number of equations → ∇2 φ = 0

Therefore we have closure. In addition, the velocity potential φ and the pressure p
are decoupled. The velocity potential φ can be solved independently first, and after
φ is obtained we can evaluate the pressure p.

p = f (v ) = f (∇φ) → Solve for φ, then find pressure.

7
(b) Bernoulli equation for P-Flow
This is a scalar equation for the pressure under the assumption of P-Flow for
steady or unsteady flow.

Euler equation:
   
∂v v2 p
+∇ − v × ω = −∇ + gy
∂t 2 ρ

Substituting v = ∇φ and ω
 = 0 into Euler’s equation above, we obtain
     
∂φ 1 p
∇ +∇ |∇φ|2 = −∇ + gy
∂t 2 ρ
or
 
∂φ 1 p
∇ + |∇φ|2 + + gy = 0,
∂t 2 ρ
which implies that

∂φ 1 p
+ |∇φ|2 + + gy = f (t)
∂t 2 ρ
everywhere in the fluid for unsteady, potential flow. The equation above can be
written as
 
∂φ 1 2
p = −ρ + |∇φ| + gy + F (t)
∂t 2
which is the Bernoulli equation for unsteady or steady potential flow.

DO NOT CONFUSE WITH

BERNOULLI EQUATION FROM § 3.4,

USED FOR STEADY, ROTATIONAL FLOW

Summary: Bernoulli equationS for ideal flow.


(a) For steady rotational or irrotational flow along streamline:
 
1 2
p = −ρ v + gy + C(ψ)
2
(b) For unsteady or steady irrotational flow everywhere in the fluid:
 
∂φ 1 2
p = −ρ + |∇φ| + gy + F (t)
∂t 2

(c) For hydrostatics, v ≡ 0, ∂t = 0:

p = −ρgy + c ← hydrostatic pressure (Archimedes’ principle)



(d) Steady and no gravity effect ( ∂t = 0, g ≡ 0):

ρv 2 ρ
p=− + c = − |∇φ|2 + c ← Venturi pressure (created by velocity)
2 2
(e) Inertial, acceleration effect:

Eulerian inertia
 
∂φ
p ∼ − ρ +···
∂t

∇p ∼ − ρ v + ···
∂t

∂p
p p+ δx
∂x
δx

9
(c) Boundary Conditions

• KBC on an impervious boundary

∂φ

v·nˆ= 
u·nˆ no flux across boundary ⇒ = Un given
∂n
n̂·∇φ Un given

• DBC: specify pressure at the boundary, i.e.,


 
∂φ 1
−ρ + |∇φ|2 + gy = given
∂t 2

Note: On a free-surface p = patm .

10

3.6.3 Stream function

• Continuity: ∇ · v = 0; Irrotationality: ∇ × v = ω = 0
• Velocity potential: v = ∇φ, then ∇ × v = ∇ × (∇φ) ≡ 0 for any φ, i.e.,
irrotationality is satisfied automatically. Required for continuity:

∇ · v = ∇2 φ = 0
 defined by

• Stream function ψ

v = ∇ × ψ 




 ≡ 0 for any ψ
Then ∇ · v = ∇ · ∇ × ψ  , i.e., satisfies continuity automatically.
Required for irrotationality:
   
 =∇ ∇·ψ
∇ × v = 0 ⇒ ∇ × ∇ × ψ  − ∇2 ψ=0 (1)



still 3 unknown
 =(ψx ,ψy ,ψz )
ψ

• For 2D and axisymmetric flows, ψ is a scalar ψ (stream functions are more ‘use­
ful’ for 2D and axisymmetric flows).


For 2D flow: v = (u, v, 0) and ∂z
≡ 0.

 
 î       
 ∂ ĵ∂ k̂∂  ∂ ∂ ∂ ∂


 
v = ∇ × ψ =
 ∂x ∂y ∂z  =
ψz î +

ψz ĵ +
ψy −
ψx k̂


ψ ψ ψ 
∂y
∂x
∂x
∂y

x y z

∂ψ
Set ψx = ψy ≡ 0 and ψz = ψ, then u = ∂y
; v = − ∂ψ
∂x

So, for 2D:

 = ∂ ψx + ∂ ψy + ∂ ψz ≡ 0
∇·ψ
∂x ∂y ∂z
Then, from the irrotationality (see (1)) ⇒ ∇2 ψ = 0 and ψ satisfies Laplace’s
equation.

11

• 2D polar coordinates: v = (vr , vθ ) and ∂z
≡ 0.

y
ê êr
r

  vr vθ
 
v
z
 êr rêθ êz       

1
 ∂ ∂  1 ∂ψ ∂ψ 1 ∂ ∂

=

v = ∇ × ψ ∂ 
=

z
êr −

z
êθ + rψθ − ψr êz
r

∂r ∂θ ∂z 
r ∂θ
∂r
r ∂r
∂θ

ψr ψθ ψz

Again let
ψr = ψθ ≡ 0 and ψz = ψ , then
1 ∂ψ ∂ψ
vr = and vθ = −
r ∂θ ∂r

 also reduces to ψ (read JNN 4.6 for details).


• For 3D but axisymmetric flows, ψ

12

• Physical Meaning of ψ.
In 2D

∂ψ ∂ψ
u= and v = −
∂y ∂x
We define
 
x  
x
ψ(x, t) = ψ(x0 , t) + v · nd

ˆ = ψ(x0 , t) + (udy − vdx)



x0 
x0
 
total volume flux
from left to right
accross a curve C
between x and x0

v
x

C’
v
t

v
xo C n̂

For ψ to be single-valued, must be path independent.


    
= or − = 0 −→ v · n
ˆ d
= ∇· v ds = 0
C C� C C� C−C � S =0, continuity

Therefore, ψ is unique because of continuity.

13

Let x1 , x2 be two points on a given streamline (v · n̂ = 0 on streamline)

streamline

x2
ψ (x2 ) = ψ (x1 ) + 
v · n̂ d

   
ψ2 ψ1 
x1 =0
along a
streamline

Therefore, ψ1 = ψ2 , i.e., ψ is a constant along any streamline. For example, on


an impervious stationary body v · n̂ = 0, so ψ = constant on the body is the
appropriate boundary condition. If the body is moving v · n̂ = Un

ψ = ψ0 + Un d
on the boddy

given

∂φ
ψ = constant ≡ =0
u=0 ∂n

ψ = given

ψo

14

Flux Δψ = −vΔx = uΔy.

∂ψ ∂ψ

Therefore, u = and v = −
∂y ∂x

(x, y + Δy)

u
streamline
-v
streamline (x +Δx, y)
(x,y)

ψ + Δψ
ψ

15

Summary of velocity potential formulation vs. stream-function formulation for ideal flows

⎧ ⎫

For irrotational flow use φ ⎬

For incompressible flow use ψ



For P-Flow use φ or ψ

velocity potential stream-function


definition v = ∇φ v = ∇ × ψ
continuity ∇ · v = 0 ∇2 φ = 0 automatically
  satisfied
 
irrotationality ∇ × v = 0 automatically satisfied ∇× ∇×ψ  =∇ ∇·ψ =0
 − ∇2 ψ


2D: w = 0, ∂z =0
continuity ∇2 φ = 0 automatically satisfied
irrotationality automatically satisfied ψ ≡ ψz : ∇2 ψ = 0

Cauchy-Riemann equations for (φ, ψ) = (real, imaginary) part of an analytic complex


function of z = x + iy

∂φ ∂ψ
u= ∂x
u= ∂y
Cartesian (x, y)
∂φ
v= ∂y v = − ∂ψ
∂x

∂φ 1 ∂ψ
vr = ∂r
vr = r ∂θ
Polar (r,θ)
1 ∂φ
vθ = r ∂θ
vθ = − ∂ψ
∂r

Given φ or ψ for 2D flow, use Cauchy-Riemann equations to find the other:

e.g. If φ = xy, then ψ = ?



∂φ ∂ψ 1 ⎪
u= =y= → ψ = y 2 + f1 (x) ⎪

∂x ∂y 2 ⎬ 1
⇒ ψ = (y 2 − x2 ) + const
∂φ ∂ψ ⎪
⎪ 2
→ ψ = − x2 + f2 (y) ⎪
1
v= =x=− ⎭

∂y ∂x 2

16
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 10

13.021 - Marine Hydrodynamics


Lecture 10

3.7 Governing Equations and Boundary Conditions for P-Flow

3.7.1 Governing Equations for P-Flow

(a) Continuity �2 φ = 0
� �
1 2
(b) Bernoulli for P-Flow (steady or unsteady) p = −ρ φt + |�φ| + gy + C(t)
2

3.7.2 Boundary Conditions for P-Flow

Types of Boundary Conditions:

∂φ
(c) Kinematic Boundary Conditions - specify the flow velocity �v at boundaries. = Un
∂n
(d) Dynamic Boundary Conditions - specify force F� or pressure p at flow boundary.
� �
1 2
p = −ρ φt + (�φ) + gy + C (t) (prescribed)
2

The boundary conditions in more detail:

• Kinematic Boundary Condition on an impermeable boundary (no flux condition)


�v ·n̂ = ����
U ·n̂ = Un = Given
���� ����
fluid velocity boundary velocity nornal boundary velocity
�v =�φ

�φ · n̂ = Un ⇒

∂ ∂ ∂
(n1 + n2 + n3 )φ = Un ⇒
∂x1 ∂x2 ∂x3

∂φ
= Un
∂n

v
U

n = (n1 , n 2 , n 3 )
v
v
( )

• Dynamic Boundary Condition: In general, pressure is prescribed

� �
1 2
p = −ρ φt + (�φ) + gy + C (t) = Given
2

3.7.3 Summary: Boundary Value Problem for P-Flow

The aforementioned governing equations with the boundary conditions formulate the

Boundary Value Problem (BVP) for P-Flow.

The general BVP for P-Flow is sketched in the following figure.

KBC : (Lecture 19)



Free surface DBC : − ρ (φ + 1 (∇φ ) 2 + gy ) + C (t ) = GIVEN
t
 21 23
 non − linear

∇ 2φ = 0

p = − ρ (φt + (∇φ ) + gy ) + C (t )
1 2

∂φ
Solid boundary KBC : = U n = GIVEN
∂n

It must be pointed out that this BVP is satisfied instantaneously.

3.8 Linear Superposition for Potential Flow


In the absence of dynamic boundary conditions, the potential flow boundary value
problem is linear.

• Potential function φ.

∇ 2 φ = 0 in V

∂φ
= U n =f on B
∂n

• Stream function ψ.

∇2 ψ = 0 in V

ψ=g on B

Linear
� Superposition: if φ1 , φ2 , . . . are harmonic functions, i.e., �2 φi = 0, then φ =
αi φi , where αi are constants, are also harmonic, and is the solution for the boundary
value problem provided the kinematic boundary conditions are satisfied, i.e.,
∂φ ∂
= (α1 φ1 + α2 φ2 + . . .) = Un on B.
∂n ∂n
The key is to combine known solution of the Laplace equation in such a way as to satisfy

the kinematic boundary conditions (KBC).

The same is true for the stream function ψ. The K.B.C specify the value of ψ on the

boundaries.

3.8.1 Example

��� �
Let φi x denote a unit-source flow with source at xi , i.e.,

��� �� � � 1
�� � �

φi x ≡ φsource x,
xi
=
ln �x −
xi
� (in 2D)

� �� � ��−1
= − 4π �x − xi � (in 3D),

then find mi such that


� �
φ= mi φi (x) satisfies KBC on B
i

Caution: φ must be regular for x ∈ V , so it is required that �x ∈


/ V.

v
•x 2

v
•x 1
∇ 2 φ = 0 in V

v
•x 3 •x 4
v

∂Φ
=f
∂n

Figure 1: Note: �xj , j = 1, . . . , 4 are not in the fluid domain V .

3.9 - Laplace equation in different coordinate systems (cf Hildebrand §6.18)


3.9.1 Cartesian (x,y,z)
� � � �
î ˆ
j ˆ
k ∂φ ∂φ ∂φ
�v = u, v, w = �φ = , ,
∂x ∂y ∂z

∂ 2φ ∂ 2φ ∂ 2φ
�2 φ = + + 2
∂x2 ∂y 2 ∂z

êz
z

P ( x, y , z )

O y ê y

x
êx

3.9.2 Cylindrical (r,θ,z)

r 2 = x2 + y 2 ,
θ = tan−1 (y/x)

� ê ê ê � � ∂φ 1 ∂φ ∂φ �
� r θ z
v = vr , vθ , vz = , ,
∂r r ∂θ ∂z

∂ 2 φ 1 ∂φ 1 ∂ 2 φ ∂ 2 φ
�2 φ = 2
+ + 2 2 + 2 ⇔
�∂r �� r ∂r� r ∂θ ∂z
r ∂r ( ∂r )
1 ∂φ
r ∂φ
� �
2 1 ∂φ ∂φ 1 ∂2φ ∂ 2φ
�φ = r + 2 2 + 2
r ∂r ∂r r ∂θ ∂z

êz
z

P (r , θ , z )

O y ê y
θ r
x
êx

3.9.3 Spherical (r,θ,ϕ)

r 2 = x2 + y 2 + z 2 ,
θ = cos−1 (z/r) ⇔ z = r (cos θ)
ϕ = tan−1 (y/x)

� � � �
� êr êθ êϕ ∂φ 1 ∂φ 1 ∂φ
v = �φ = vr , vθ , vϕ = , ,
∂r r ∂θ r(sin θ) ∂ϕ

� �
2 ∂ 2 φ 2 ∂φ 1 ∂ ∂φ 1 ∂ 2φ
�φ = + + sin θ + ⇔
2 2
�∂r �� r ∂r� r sin θ ∂θ ∂θ r2 sin2 θ ∂ϕ2
1 ∂
r 2 ∂r
(r2 ∂φ
∂r )
� � � �
2 1 ∂ 2 ∂φ 1 ∂ ∂φ 1 ∂ 2φ
�φ = 2 r + 2 sin θ + 2 2
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2

êz
z

P (r , θ , φ )

θ
r

O y ê y
φ
x
êx

3.10 Simple Potential flows


1. Uniform Stream �2 (ax + by + cz + d) = 0


1D: φ = U x + constant ψ = U y +
constant;
v = (U, 0, 0)

2D: φ = U x + V y + constant ψ = U y − V x +
constant;
v = (U, V, 0)

3D: φ = U x + V y + W z +
constant
v = (U, V, W )

2. Source (sink) flow

2D, Polar coordinates


� � �
2 1 ∂ ∂ 1 ∂2
� = r + 2 2 , with r = x2 + y 2
r ∂r ∂r r ∂θ
An axisymmetric
� solution: φ = a ln r + b. Verify that it satisfies �2 φ = 0, except at
r = x2 + y 2 = 0. Therefor, r = 0 must be excluded from the flow.

Define 2D source of strength m at r= 0:

m
φ= ln r

∂φ m m
�φ = êr = êr ⇐⇒ vr = , vθ = 0
∂r 2πr 2πr

source
y
(strength m)

Net outward volume flux is

� �
�� �
�� �
v · n̂ds = � · vds = � · vds
C S Sε
� �
�2π
v · n̂ds = vr rε dθ = m
����
Cε 0
����
m
2πrε source
strength
y
n̂ C

x
ε

If m < 0 ⇒ sink. Source m at (x0 , y0 ):


m
φ = ln (x − x0 )2 + (y − y0 )2

m m
φ = ln r (Potential function) ←→ ψ = θ (Stream function)
2π 2π

m
Ψ= θ
y 2π
θ
m
Vr =

x
1
ψ=0

10

3D: Spherical coordinates


� � � � �
2 1 ∂ 2 ∂ ∂ ∂
� = 2 r + , , · · · , where r = x2 + y 2 + z 2
r ∂r ∂r ∂θ ∂ϕ
a
A spherically symmetric solution: φ = + b. Verify �2 φ = 0 except at r = 0.
r

Define a 3D source of strength m at r = 0. Then

m ∂φ m
φ=− ⇐⇒ vr = = , vθ = 0, vϕ = 0
4πr ∂r 4πr2
Net outward volume flux is
��
m
� vr dS = 4πrε2 · = m (m < 0 for a sink )
4πrε2

11

3. 2D point vortex
� �
2 1 ∂ ∂ 1 ∂2
� = r + 2 2
r ∂r ∂r r ∂θ
Another particular solution: φ = aθ + b. Verify that �2 φ = 0 except at r = 0.

Define the potential for a point vortex of circulation Γ at r = 0. Then

Γ ∂φ 1 ∂φ Γ
φ= θ ⇐⇒ vr = = 0, vθ = = and,
2π ∂r r ∂θ 2πr
1 ∂
ωz = (rvθ ) = 0 except at r = 0
r ∂r
Stream function:

Γ
ψ=− ln r

Circulation:

� � � �2π
� � � � � � Γ
v · dx = v · dx + v · dx = rdθ = Γ
����
2πr
C1 C2 C1 −C2
� R R �� �
0 vortex
ωz dS=0 strength
S

12

4. Dipole (doublet flow)

A dipole is a superposition of a sink and a source with the same strength.

2D dipole:

� � � �
m 2 2 2 2
φ= ln (x − a) + y − ln (x + a) + y

� �
µ ∂ �
lim φ = ln (x − ξ) + y ��
2 2
a→0 2π
���� ∂ξ ξ=0
µ = 2ma
constant
µ x µ x
=− = −
2π x2 + y 2 2π r2

2D dipole (doublet) of moment µ at the origin oriented in the +x direction.


NOTE: dipole = µ ∂ξ (unit source)

13

unit
source α
x

−µ x cos α + y sin α −µ cos θ cos α + sin θ sin α


φ= 2 2
=
2π x +y 2π r
3D dipole:

⎛ ⎞
m

1 1

where µ = 2ma fixed

φ = lim

� −

a→0 4π
2 2 2 2 2 2
(x − a)
+ y + z (x + a)
+ y + z


µ ∂ 1
� µ x µ x

= −
� � = − = −
4π ∂ξ
� 4π (x2 + y 2 + z 2 )3/2 4π r3
(x
− ξ)2 + y 2 + z 2 �
ξ=0

3D dipole (doublet) of moment µ at the origin oriented in the +x direction.

14

5. Stream and source: Rankine half-body

It is the superposition of a uniform stream of constant speed U and a source of


strength m.

U
m

m � 2
2D: φ = U x + ln x + y 2

U
m
x D

v
stagnation point v = 0

Dividing
Streamline

∂φ m x
u= =U+
∂x 2π x + y 2
2
m
u|y=0 = U + , v |y=0 = 0 ⇒
2πx
m
V� = (u, v) = 0 at x = xs = − , y=0
2πU

m
For large x, u → U , and U D = m by continuity ⇒ D = .
U

15

m
3D: φ = U x − �
4π x
2 + y 2 + z 2

stagnation point

div. streamlines

∂φ m x
u= =U+
∂x 4π (x + y + z 2 )3/2
2 2
m x
u|y=z=0 = U + , v |y=z=0 = 0, w|y=z=0 = 0 ⇒
4π |x|3

� m
V = (u, v, w) = 0 at x = xs = − , y=z=0
4πU

m
For large x, u → U and U A = m by continuity ⇒ A = .
U

16

6. Stream + source/sink pair: Rankine closed bodies

U S +m -m S x
a

dividing streamline
(see this with PFLOW)


To have a closed body, a necessary condition is to have min body = 0

2D Rankine ovoid:

� � � � � �
m 2 2 2 2
m (x + a)2 + y 2
φ = U x+ ln (x + a) + y − ln (x − a) + y = U x+ ln
2π 4π (x − a)2 + y 2

3D Rankine ovoid:
⎡ ⎤
m⎣ 1 1

φ = Ux − � −�
4π 2 2
(x + a) + y2 + z2 (x − a) + y2 + z2

17

For Rankine Ovoid,

� �
∂φ m x+a x−a
u= =U+ −
∂x 4π �(x + a)2 + y 2 + z 2 �3/2 �(x − a)2 + y 2 + z 2 �3/2
� �
m 1 1
u|y=z=0 =U + −
4π (x + a)2 (x − a)
2
m (−4ax)

=U +
4π (x2 − a2 )2

� �2 � m

u|y=z=0 =0 at x2 − a2 = 4ax
4πU

At x = 0,

m 2a
u=U+ where R = y 2 + z 2
4π (a2 + R2 )3/2

Determine radius of body R0 :

�R0
2π uRdR = m
0

18

7. Stream + Dipole: circles and spheres

r
U µ θ

µx � µ �
2D: φ = U x + = cos θ U r +
2πr2 ↑ 2πr
x=r cos θ
The radial velocity is then
∂φ � µ �
ur = = cos θ U − .
∂r 2πr2
� µ
Setting the radial velocity vr = 0 on r = a we obtain a = 2πU
. This is the K.B.C.
for a stationary circle of radius a. Therefore, for

µ = 2πU a2

the potential � µ �
φ = cos θ U r +
2πr
is the solution to ideal flow past a circle of radius a.
• Flow past a circle (U, a).

19

� �
a2
φ = U cos θ r + r
� 2

1 ∂φ
Vθ = r ∂θ
= −U sin θ 1 + ar2

= 0 at θ = 0, π − stagnation points
Vθ |r=a = −2U sin θ π 3π
= �2U at θ = 2 , 2 − maximum tangential velocity

2U

2U

Illustration of the points where the flow reaches maximum speed around the circle.

µ cos θ � µ �
3D: φ = U x + = U r cos θ 1 +
4π r2 4πr3

y
µ
r
U θ
x
z

The radial velocity is then


∂φ � µ �
vr = = cos θ U −
∂r 2πr3

20
� µ
Setting the radial velocity vr = 0 on r = a we obtain a = 3
2πU
. This is the K.B.C.
for a stationary sphere of radius a. Therefore, choosing

µ = 2πU a3

the potential � µ �
φ = cos θ U r +
2πr
is the solution to ideal flow past a sphere of radius a.
• Flow past a sphere (U, a).

� �
a3
φ = Ur cos θ 1 + 3
2r
� �
1 ∂φ a3
vθ = = −U sin θ 1 + 3
r ∂θ 2r

3U = 0 at θ = 0, π
vθ |r=a = − sin θ
2 = − 3U
2
at θ = π2

3/ U
2

θ
x

3/ U
2

21

8. 2D corner flow Velocity potential φ = rα cos αθ; Stream function ψ = rα sin αθ


� 2 �
∂ 1 ∂ 1 ∂2
(a) �2 φ = ∂r 2 + r ∂r
+ r2 ∂θ2
φ=0
(b)

∂φ
ur = = αrα−1 cos αθ
∂r

1 ∂φ

uθ = = −αrα−1 sin αθ
r ∂θ
∴ uθ = 0 { or ψ = 0} on αθ = nπ, n = 0, ±1, ±2, . . .

i.e., on θ = θ0 = 0, απ , 2π
α
, . . . (θ0 ≤ 2π)

i. Interior corner flow – stagnation point origin: α > 1. For example,

α = 1, θ0 = 0, π, 2π, u = 1, v = 0

ψ=0

22

π 3π
α = 2, θ0 = 0, , π, ,2π u = 2x, v = −2 y
o 2 2
(90 corner)

ψ=0

ψ=0

θ=2π/3, ψ = 0

2π 4π θ=0, ψ = 0
α = 3 2 , θ0 = 0, , ,2π 120o 120o
3 3 θ=2π, ψ = 0
o
120
(120o corner)

θ=4π/3, ψ = 0

23

ii. Exterior corner flow, |v| → ∞ at origin:

α<1
π
θ0 = 0, only
α
π
Since we need θ0 ≤ 2π, we therefore require α
≤ 2π, i.e., α ≥ 1/2 only.

1/2 ≤ α < 1
π
θ0 = 0,
α
For example,

α = 1/2, θ0 = 0, 2π (1/2 infinite plate, flow around a tip)

θ=0, ψ = 0

θ=2π, ψ = 0

α = 2/3, θ0 = 0, 3π
2
(90o exterior corner)

θ=0, ψ = 0

θ=3π/2, ψ = 0

24

Appendix A1: Summary of Simple Potential Flows

Cartesian Coordinate System


Flow Streamlines Potential Stream function
φ(x, y, z) ψ(x, y)

Uniform flow U∞ x + V∞ y + W∞ z U∞ y − V∞ x

y−yo
2D Source/Sink (m) at (xo , yo ) m

ln((x − xo )2 + (y − yo )2 ) m

arctan( x−x )
o

3D Source/Sink (m) at (xo , yo , zo ) m q


− 4π 1 NA
(x−xo )2 +(y−yo )2 +(z−zo )2

Γ y−yo Γ ln((x − x )2 + (y − y )2 )
Vortex (Γ) at (xo , yo ) 2π
arctan( x−x ) − 2π o o
o

µ (x−xo ) cos α+(y−yo ) sin α µ (y−yo ) cos α+(x−xo ) sin α


2D Dipole (µ) at (xo , yo ) at an angle α − 2π
(x−xo )2 +(y−yo )2 2π (x−xo )2 +(y−yo )2
α

µ (x−xo )
3D Dipole (+x) (µ) at (xo , y0 , zo ) − 4π NA
((x−xo )2 +(y−yo )2 +(z−zo )2 )3/2

25

Appendix A2: Summary of Simple Potential Flows

Cylindrical Coordinate System


Flow Streamlines Potential Stream function
φ(r, θ, z) ψ(r, θ)

Uniform flow U∞ r cos θ + V∞ r sin θ + W∞ z U∞ r sin θ − V∞ r cos θ

2D Source/Sink (m) at (xo , yo ) m ln r m θ


2π 2π

3D Source/Sink (m) at (xo , yo , zo ) m


− 4πr NA

Vortex (Γ) at (xo , yo ) Γ θ Γ ln r


− 2π

µ cos θ cos α+sin θ sin α µ sin θ cos α+cos θ sin α


2D Dipole (µ) at (xo , yo ) at an angle α − 2π r 2π r
α

µ cos θ
3D Dipole (+x) (µ) at (xo , yo , zo ) − 4π 2 NA
r

26

Appendix A3: Combination of Simple Potential Flows

m m m
Stream + Source (2D) φ = U∞ x + 2π ln r xs = − 2πU ∞
D= U∞

=

m √ 1 m m
Rankine Half Body (3D) φ = U∞ x − 4π xs = − 4πU∞ A= U∞
x2 +y 2 +z 2

m
� �
Stream + Source + Sink (2D) φ = U∞ x + 2π ln((x + a)2 + y 2 ) − ln((x − a)2 + y 2 )

=
m √ 1 1
Rankine Closed Body (3D) φ = U∞ x + 4π ( (x+a)2 +y 2 +z 2 −√ )
(x−a)2 +y 2 +z 2

µx a2
Stream + Dipole (2D) φ = U∞ x + 2πr 2 if µ = 2πa2 U∞ φ = U∞ cos θ(r + r )

=
µ cos θ a3
Circle (Sphere) R = a (3D) φ = U∞ x + 4πr 2 if µ = 2πa3 U∞ φ = U∞ cos θ(r + 2r 2 )

2D Corner Flow (2D) φ = Crα cos(αθ) ψ = Crα sin(αθ) θ0 = 0, nπ


α

27

Appendix B: Far Field Behavior of Simple Potential Flows

Far field behavior


φ �v = �φ
r >> 1

1
(2D) ∼ ln r ∼
r
Source
1 1
(3D) ∼ ∼
r r2

1 1
(2D) ∼ ∼
r r2
Dipole
1 1
(3D) ∼ ∼
r2 r3

1
Vortex (2D) ∼1 ∼
r

28

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 11

13.021 - Marine Hydrodynamics


Lecture 11

3.11 - Method of Images


m  2
• Potential for single source: φ = ln x + y 2

m
m

   
m 2 2
• Potential for source near a wall: φ = ln x2 + (y − b) + ln x2 + (y + b)

m
m
y

dφ b
=0 x
dy

b
m

Added source for


symmetry

Note: Be sure to verify that the boundary conditions are satisfied by symmetry or
by calculus for φ (y) = φ (−y).

 
Γ −1 y − b −1 y + b
• Vortex near a wall (ground effect): φ = Ux + tan ( ) − tan ( )
2π x x

U b y

x
b

Added vortex for


symmetry


Verify that = 0 on the wall y = 0.
dy
 
a2 a2
• Circle of radius a near a wall: φ ∼
= Ux 1 + +
x2 + (y − b)2 x2 + (y + b)2

y
U b
y
x

∂φ
This solution satisfies the boundary condition on the wall ( ∂n = 0), and the degree it
satisfies the boundary condition of no flux through the circle boundary increases as
the ratio b/a >> 1, i.e., the velocity due to the image dipole small on the real circle
for b >> a. For a 2D dipole, φ ∼ d1 , ∇φ ∼ d12 .

2
• More than one wall:

Example 1:
b
b
U
U
b'
b'

b'

Example 2: Example 3:
b b
b b
-Γ Γ
b' b'
b' b'

b' b'
b' b'
Γ -Γ
b b b b

3.12 Forces on a body undergoing steady translation


“D’Alembert’s paradox”
3.12.1 Fixed bodies & translating bodies - Galilean transformation.
y y’

x’ U
x o’
o z’
z

Fixed in space Fixed in translating body

x = x` + Ut

Reference system O: v , φ, p Reference system O’: v  , φ , p

X S X’ S U
O O’

∇2 φ = 0 ∇2 φ = 0

v · n̂ = ∂φ  · n̂ = (U, 0, 0) · (nx , ny , nz )
=U
∂n  ∂φ
v · n̂ = =0
= U nx on Body ∂n

 
v → 0 as |x| → ∞ v → (−U, 0, 0) as |x | → ∞
φ → 0 as |x| → ∞ φ → −U x as |x | → ∞
Galilean transform:
 
v(x, y, z, t) = v (x = x − U t, y, z, t) + (U, 0, 0)
φ(x, y, z, t) = φ (x = x − U t, y, z, t) + U x ⇒
−U x + φ(x = x + U t, y, z, t) = φ (x , y, z, t)
Pressure (no gravity)
p∞ = − 12 ρv 2 + Co = Co = − 12 ρv 2 + Co = Co − 12 ρU 2

∴ Co = Co − 12 ρU 2
In O: unsteady flow In O’: steady flow
∂φ 1
ps = −ρ ∂φ
∂t
v 2 +Co
− 12 ρ  ps = −ρ v 2 +Co = Co
− 2 ρ 
∂t

U2 0
0
∂φ ∂ ∂x ∂
∂t
=( + ∂x
) (φ + U x ) = −U 2
∂t ∂t
 
0 −U

∴ ps = ρU 2
− 12 ρU 2 + Co = 12 ρU 2 + Co

ps − p∞ = 12 ρU 2 stagnation pressure ps − p∞ = 12 ρU 2 stagnation pressure

3.12.2 Forces


Total fluid force for ideal flow (i.e., no shear stresses): F =  pˆ
ndS
B
For potential flow, substitute for p from Bernoulli:
⎛ ⎞

⎜ ∂φ 1 ⎟
 ⎜ ⎟
F =  −ρ ⎜ + |∇φ|2 + gy +c(t)⎟ ndS
ˆ
⎝∂t 
2  ⎠
B
hydrodynamic hydrostatic
force force
 
For the hydrostatic case v ≡ φ ≡ 0 :


F s =  (−ρgyn̂) dS = (−) ∇ (−ρgy) dυ = ρg∀ĵ where ∀ = dυ
↑ ↑  
Gauss outward
B theorem normal
υB Archimedes υB
principle

We evaluate only the hydrodynamic force:


 
 ∂φ 1 2
F d = −ρ + |∇φ| n̂dS
∂t 2
B
 ∂φ 
For steady motion ∂t
≡0 :

 1
Fd = − ρ v 2 ndS
ˆ
2
B

5
3.12.3 Example Hydrodynamic force on 2D cylinder in a steady uniform stream.


U S x
a

   −ρ  2π 
−ρ 

F d =
2
|∇φ|2 nd
ˆ = 2
2
|∇φ|
r=a n̂adθ
B 0
 2π
Fx =
F ·
î = −ρa
2
dθ |∇φ|
2r=a 

·

0
− cos θ
2π 2
=
ρa
2
|∇φ|
r=a cos θdθ

Velocity potential for flow past a 2D cylinder:




a2
φ = U r cos θ 1 + 2
r
Velocity vector on the 2D cylinder surface:


  

∂φ
 1 ∂φ

∇φ|r=a = ( vr |r=a , vθ |r=a ) =  ,

∂r


r ∂θ





r=a

0 r=a −2U sin θ

Square of the velocity vector on the 2D cylinder surface:



|∇φ|2 
r=a = 4U 2 sin2 θ

6
Finally, the hydrodynamic force on the 2D cylinder is given by




ρa   1 
Fx = dθ 4U 2 sin2 θ cos θ = ρU 2 (2a) 2 dθ sin 2
  θ cos
 θ = 0
2
0
2   0 even odd
π 3π
diameter w.r.t ,
ps −p∞ or
projection  
2 2

≡0

Therefore, Fx = 0 ⇒ no horizontal force ( symmetry fore-aft of the streamlines). Similarly,



1 
Fy = ρU 2 (2a)2 dθ sin2 θ sin θ = 0
2
0

In fact, in general we find that F ≡ 0, on any 2D or 3D body.

D’Alembert’s “paradox”:

No hydrodynamic force∗ acts on a body moving with steady translational (no circulation)
velocity in an infinite, inviscid, irrotational fluid.


The moment as measured in a local frame is not necessarily zero.

3.13 Lift due to Circulation


3.13.1 Example Hydrodynamic force on a vortex in a uniform stream.
Γ Γ
φ = Ux + θ = U r cos θ + θ
2π 2π

U Γ

Consider a control surface in the form of a circle of radius r centered at the point vortex.
Then according to Newton’s law:
d  steady flow
ΣF = LCV −→
dt
(FV + FCS ) + M
 N ET = 0 ⇔ F ≡ −FV = FCS + M
 N ET

Where,
F = Hydrodynamic force exerted on the vortex from the fluid.
FV = −F = Hydrodynamic force exerted on the fluid in the control volume from the vortex.
FCS = Surface force (i.e., pressure) on the fluid control surface.
M N ET = Net linear momentum flux in the control volume through the control surface.
d 
L
dt CV
= Rate of change of the total linear momentum in the control volume.

Γ Fy θ
U
Fx
x

Control
volume

The hydrodynamic force on the vortex is F = FCS + M


 IN

8
a. Net linear momentum flux in the control volume through the control surfaces, M  N ET .
Recall that the control surface has the form of a circle of radius r centered at the point
vortex.

a.1 The velocity components on the control surface are

u = U− sin θ
2πr
Γ
v = cos θ
2πr

The radial velocity on the control surface is therefore, given by

∂x 
ur = U = U cos θ = V · n̂
∂r

Γ
vθ =
2πr
U
θ

a.2 The net horizontal and vertical momentum fluxes through the control surface are
given by



2π  
Γ
(MN ET )x = − ρ dθruvr = − ρ dθr U − sin θ U cos θ = 0
2πr
0 0


2π  
Γ
(M N ET )y = − ρ dθrvvr = − ρ dθr cos θ U cos θ
2πr
0 0


ρU Γ ρU Γ
=− cos2 θdθ = −
2π 2
0

9
b. Pressure force on the control surface, FCS .

b.1 From Bernoulli, the pressure on the control surface is


1
p = − ρ |v |2 + C
2
b.2 The velocity |v |2 on the control surface is given by
 2  2
2 2 2 Γ Γ
|v | =u + v = U − sin θ + cos θ
2πr 2πr
 2
2 Γ Γ
=U − U sin θ +
πr 2πr

b.3 Integrate the pressure along the control surface to obtain FCS

2π

(FCS )x = dθrp(− cos θ) = 0


0


2π  ρ
 
(FCS )y = dθrp(− sin θ) = − 2 ΓU
− πr (−r) dθ sin2 θ = − 12 ρU Γ
0
0 
π

c. Finally, the force on the vortex F is given by

Fx = (FCS )x + (Mx )IN = 0


Fy = (FCS )y + (My )IN = −ρU Γ
i.e., the fluid exerts a downward force F = −ρU Γ on the vortex.
Kutta-Joukowski Law

2D : F = −ρU Γ
3D : F = ρU
 × Γ
Generalized Kutta-Joukowski Law:
 

n
F = ρU
 × Γi
i=1

where F is the total force on a system of n vortices in a free stream with speed U
.

10
13.021 – Marine Hydrodynamics

Lecture 12

13.021 - Marine Hydrodynamics


Lecture 12

3.14 Lifting Surfaces


3.14.1 2D Symmetric Streamlined Body

No separation, even for large Reynolds numbers.

stream line

• Viscous effects only in a thin boundary layer.


• Small Drag (only skin friction).
• No Lift.

3.14.2 Asymmetric Body

(a) Angle of attack α,

chord line

U α

(b) or camber η(x),

chord line mean camber line

(c) or both

amount of camber
chord line
mean camber line

U α

angle of attack

 and Drag  to U
Lift ⊥ to U 

3.15 Potential Flow and Kutta Condition


From the P-Flow solution for flow past a body we obtain

P-Flow solution, infinite velocity at trailing edge.


Note that (a) the solution is not unique - we can always superimpose a circulatory flow
without violating the boundary conditions, and (b) the velocity at the trailing edge →
∞. We must therefore, impose the Kutta condition, which states that the ‘flow leaves
tangentially the trailing edge, i.e., the velocity at the trailing edge is finite’.
To satisfy the Kutta condition we need to add circulation.

Circulatory flow only.

Superimposing the P-Flow solution plus circulatory flow, we obtain

Figure 1: P-Flow solution plus circulatory flow.

3.15.1 Why Kutta condition?

Consider a control volume as illustrated below. At t = 0, the foil is at rest (top control
volume). It starts moving impulsively with speed U (middle control volume). At t = 0+ ,
a starting vortex is created due to flow separation at the trailing edge. As the foil moves,
viscous effects streamline the flow at the trailing edge (no separation for later t), and the
starting vortex is left in the wake (bottom control volume).

t=0 Γ=0

ΓS
+ ΓS
t=0 U starting vortex
due to separation
(a real fluid effect,
no infinite vel of
potetial flow)

ΓS
for later
ΓS
U
t

no Γ

starting vortex left in wake

Kelvin’s theorem:

= 0 → Γ = 0 for t ≥ 0 if Γ(t = 0) = 0
dt
After a while the ΓS in the wake is far behind and we recover Figure 1.

4
3.15.2 How much ΓS ?

Just enough so that the Kutta condition is satisfied, so that no separation occurs. For

example, consider a flat plate of chord  and angle of attack α, as shown in the figure

below.

chord length

Simple P-Flow solution

Γ = πlU sin α
L = ρU Γ = ρU 2 πl sin α
|
|L
CL = 1 2 = 2π sin α ≈ 2πα for small α
ρU l � �� �
2 only for
small α

However, notice that as α increases, separation occurs close to the leading edge.

Excessive angle of attack leads to separation at the leading edge.

When the angle of attack exceeds a certain value (depends on the wing geometry) stall
occurs. The effects of stalling on the lift coefficient (CL = 1 ρU 2 Lspan ) are shown in the
2
following figure.

C
L

This region independent of R,

ν used only to get Kutta

condition

stall location f(R)

stall

α
O(5 o )

• In experiments, CL < 2πα for 3D foil - finite aspect ratio (finite span).

• With sharp leading edge, separation/stall to early.

sharp trailing edge


round leading edge to forstall to develop circulation
stalling

3.16 Thin Wing, Small Angle of Attack


• Assumptions

– Flow: Steady, P-Flow.


– Wing: Let yU (x), yL (x) denote the upper and lower vertical camber coordinates,
respectively. Also, let x = /2, x = −/2 denote the horizontal coordinates of
the leading and trailing edge, respectively, as shown in the figure below.

y=yU(x)

For thin wing, at a small angle of attack it is

yU yL
, << 1
 
dyU dyL
, << 1
dx dx
The problem is then linear and superposition applies.
Let η(x) denote the camber line
1 t(x)
η(x) = (yU (x) + yL (x)),
2
and t(x) denote the half-thickness
t(x)
Camber line η(x)
1
t(x) = (yU (x) − yL (x)).
2

For linearized theory, i.e. thin wing at small AoA, the lift on the wing depends
only on the camber line but not on the wing thickness. Therefore, for the
following analysis we approximate the wing by the camber line only and ignore
the wing thickness.

7
• Definitions

In general, the lift on the wing is due to the total circulation Γ around the wing.
This total circulation can be given in terms due to a distribution of circulation γ(x)
(Units: [LT −1 ]) inside the wing, i.e.,
� /2
Γ= γ(x)dx
−/2

γ (x) Γ
U

Noting that superposition applies, let the total potential Φ for this flow be expressed
as the sum of two potentials

Φ = −U
� �� x� + φ
����
Free stream Disturbunce
potential potential

The flow velocity can by expressed as

v = ∇Φ = (−U + u, v)

where (u, v) are given by ∇φ = (u, v) and denote the velocity disturbance, due to the
presence of the wing. For linearized wing we can assume
u v
u, v << U ⇒ , << 1
U U

Consider a flow property q, such as velocity, pressure etc. Then let qU = q(x, 0+ ) and
qL = q(x, 0− ) denote the values of q at the upper and lower wing surfaces, respectively.

• Lift due to circulation

Applying Bernoulli equation for steady, inviscid, rotational flow, along a streamline
from ∞ to a point on the wing, we obtain
1 � �
p − p∞ = − ρ |v |2 − U 2 ⇒
2
1 �� � � 1
p − p∞ = − ρ (u − U )2 + v 2 − U 2 = − ρ(u2 + v 2 − 2uU ) ⇒
2 2
1 u v v
p − p∞ = − ρuU ( + −2)
2 U
���� U ����
���� u
<<1 <<1 ∼1

Dropping terms of order Uu , v


U
<< 1 we obtained the linearized Bernoulli equation
for thin wing at small AoA
p − p∞ = ρuU

Integrating the pressure along the wing surface, we obtain an expression for the total
lift L on the wing

� �l/2
�� � � ��
L = (p − p∞ )ny dS = p(x, 0− ) − p∞ − p(x, 0+ ) − p∞ dx
−l/2

�l/2 �l/2
� � � �
L = p(x, 0− ) − p(x, 0+ ) dx = ρU u(x, 0− ) − u(x, 0+ ) dx (1)
−l/2 −l/2

To obtain the total lift on the wing we will seek an expression for u(x, 0± ).

Consider a closed contour on the wing, of negligible thickness, as shown in the figure

below. γ (x)

u ( x,0 + )
x
t→0
u ( x,0 − )

δx

In this case we have


γ(x)δx = |u(x, 0+ )|δx + u(x, 0− )δx ⇒ γ(x) = |u(x, 0+ )| + u(x, 0− )

For small u/U we can argue that u(x, 0+ ) ∼


= −u(x, 0− ), and obtain
γ(x)
u(x, 0± ) = ∓ (2)
2
From Equations (1), and (2) the total lift can be expressed as
� l/2
L = ρU γ(x)dx = ρU Γ
−l/2
� �� �

The same result can be obtained from the Kutta-Joukowski law (for nonlinear foil)
� /2
δL = ρU δΓ = ρU γ(x)δx ⇒ L = ρU γ(x)δx = ρU Γ
−/2

δ L = ρU δ Γ = ρUγ (x)δ x

x U
t→0
δ Γ = γ (x)δ x

δx

10

• Moment, with respect to mid-chord, due to circulation

y
L

l xcp l x
2 2
M

δL(x) = ρU γ(x)δx
δM = xδL(x) = ρU xγ(x)δx ⇒
� /2
M = ρU xγ(x)dx ⇒
−/2
M
CM = 1
2
ρU 2 2

The center of pressure xcp , can be obtained by

M = Lxcp ⇒
� /2
M −/2
xγ(x)dx
xcp = = � /2
L γ(x)dx
−/2

11

� /2
3.17 Simple Closed-Form Solutions for −/2 γ(x)dx from Linear
Theory
1. Flat plate at angle of attack α, i.e., η = αx.

Linear lifting theory gives γ(x), which can be integrated to give the lift coefficient

CL ,
� /2
L/span = ρU γ(x)dx = · · · = ρU 2 πα ⇒
−/2
L/span
CL = 1 ⇒
2
ρU 2 
CL = 2πα ( exact nonlinear hydrofoil CL = 2π sin α)

the moment coefficient CM ,


� /2
M/span = ρU xγ(x)dx = · · · = 14 ρU 2 2 πα ⇒
−/2
M/span
CM = 1 ⇒
2
ρU 2 2
1
CM = 2
πα

and the center of pressure xcp

xcp = 14  i.e., at quarter chord

12

2. Parabolic camber η = η0 {1 − ( 2xl )2 }, at zero AoA α = 0.

Linear lifting theory gives γ(x), which can be integrated to give the lift coefficient

CL ,
� /2
L/span = ρU γ(x)dx = · · · = 2ρU 2 πη0 ⇒
−/2
η0 η0
CL = 4π , where ≡ ‘camber ratio’
 

the moment coefficient CM ,

M/span = 0 (from symmetry) ⇒


CM = 0

and the center of pressure xcp


xcp = 0

13

� � �2 �
2x
3. Linear superposition: Both AoA and camber η = αx + η0 1 − .


η0
CL = CLα + CLη = 2πα + 4π

We can also write the previous relation in a more general form

CL (α) = 2πα + CL (α = 0)
� �� �
≡ 4π ηl0

η0
Lift coefficient CL as a function of the angle of attack α and l
.

In practice even if the camber is not parabolic, we still make use of the
previous relations, i.e., CL (α = 0) ∼= 4πη0 /.
Also note that the angle of attack for any camber is defined as
η(/2) − η(−/2) yU − yL
α≡ =
 
and η0 is determined from η ∗ , where

η ∗ = η − αx.

14
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 13

13.021 - Marine Hydrodynamics


Lecture 13

3.18 Unsteady Motion - Added Mass


D’Alembert: ideal, irrotational, unbounded, steady.

Example Force on a sphere accelerating (U = U (t), unsteady) in an unbounded fluid that


is at at rest at infinity.

ϕ θ
r U(t)

x
3D Dipole
a

U(t)


∂φ

K.B.C on sphere:
= U (t) cos θ

∂r

r=a
Solution: Simply a 3D dipole (no stream)

a3
φ = −U (t) cos θ
2r2

∂φ 
Check: = U (t) cos θ

∂r

r=a

Hydrodynamic force:
  
∂φ 1 2
Fx = −ρ + |∇φ| nx dS
∂t 2
B

On r = a,


∂φ  3
˙ a cos θ| 1˙
 = − U r=a = − U a cos θ
∂t r=a 2r 2 2
   
∂φ 1 ∂φ 1 ∂φ 1
∇φ|r=a = , , = U cos θ, U sin θ, 0
∂r r ∂θ r sin θ ∂ϕ 2
 1
|∇φ|2 r=a = U 2 cos2 θ + U 2 sin2 θ; n̂ = −êr , nx = − cos θ
4
 π
dS = (adθ) (2πa sin θ)
B 0

adθ
a sin θ
a
θ
x

Finally, ⎡ ⎞⎤ ⎛
⎛ ⎞

π ⎢ 1 ⎜ 1 2 2 ⎟ ⎥
2 ⎝ ⎠ ⎢ ˙ 1 ⎜ 2 2 ⎟ ⎥
Fx = (−ρ) 2πa dθ (sin θ) − cos
 θ ⎢− U a cos θ + 2 ⎝U cos θ + U sin θ⎠⎥
0 ⎣ 2    4  ⎦
nx
∂φ
∂t
|∇φ|2
π π 
3 2 2 2 2 1 2
Fx = −U̇ (ρa )π dθ sin θ cos θ + (ρU )πa dθ sin θ cos θ cos θ + sin θ
4
0   0
  

2/3 ⎤ = 0, D’alembert revisited

⎢ 2 3 ⎥

Fx ⎢
= − U̇ (t) ⎣ ρ πa ⎥

   ⎦


Hydrodynamic Force Acceleration Fluid Density

3  
Volume =1/2∀sphere

Thus the Hydrodynamic Force on a sphere of diameter a moving with velocity U (t) in
an unbounded fluid of density ρ is given by
 
2 3
Fx = −U̇ (t) ρ πa
3

Comments:

• If U̇ = 0 → Fx = 0, i.e., steady translation → no force (D’Alembert’s Condition ok).

• Fx ∝ U̇ with a (−) sign, i.e., the fluid tends to ‘resist’ the acceleration.

• [· · · ] has the units of (fluid) mass ≡ ma

• Equation of Motion for a body of mass M that moves with velocity U :


 
M
 U̇ = ΣF = F H + FB = − U̇ m a + FB ⇔
  
Body mass Hydrodynamic force All other forces on body Fluid mass
S pn̂dS

(M + ma ) U̇ = FB

i.e., the presence of fluid around the body acts as an added or virtual mass to the
body.

3
3.19 General 6 Degrees of Freedom Motions
3.19.1 Notation Review
(3D) U1 , U2 , U3 : Translational velocities
U4 ≡ Ω1 , U5 ≡ Ω2 , U6 ≡ Ω3 : Rotational velocities

2
5
1
6 4

(2D) U1 , U2 : Translational velocities


U6 ≡ Ω3 : Rotational velocity
U3 = U 4 = U 5 = 0

2
6
1

3.19.2 Added Mass Tensor (matrix)

mij ; i, j = 1, 2, 3, 4, 5, 6

mij : associated with force on body in i direction due to unit acceleration in j


direction. For example, for a sphere:

m11 = m22 = m33 = 1/2ρ∀ = (mA ) all other mij = 0

3.19.3 Added Masses of Simple 2D Geometries

• Circle
2

a
1

m11 = m22 = ρ∀ = ρπa2

• Ellipse

a
1
b

m11 = ρπa2 , m22 = ρπb2

• Plate

1
2a

m11 = ρπa2 , m22 = 0

• Square

2a
1

2a

m11 = m22 ≈ 4.754ρa2

A reasonable approximation to estimate the added mass of a 2D body is to


use the displaced mass (ρ∀) of an ‘equivalent cylinder’ of the same lateral
dimension or one that ‘rounds off’ the body. For example, consider a square
and approximate with an
(a) inscribed circle: mA = ρπa2 = 3.14ρa2 .

2a

√ 2
(b) circumscribed circle: mA = ρπ 2a = 6.28ρa2 .

(√2)a

Arithmetic mean of (a) + (b) ≈ 4.71ρa2 .

3.19.4 Generalized Forces and Moments


In this paragraph we are looking at the most general case where forces and moments
are induced on rigid body moving with 6 DoF motions, in an unbounded fluid that
is at rest at infinity.

Body fixed reference frame, i.e., OX1 X2 X3 is fixed on the body.

x2
v
U( t )

o x1

x3
v
Ω( t )

 (t) = (U1 , U2 , U3 ) , translational velocity

U

= (Ω1 , Ω2 , Ω3 ) ≡ (U4 , U5 , U6 ) , rotational velocity with respect to O
Ω(t)

 , Ω),
Consider a body with a 6 DoF motion (U  and a fixed reference frame OX1 X2 X3 .
Then the hydrodynamic forces and moments with respect to O are given by the
following relations (JNN §4.13)

• Forces

Fj = −U̇i mji − Ejkl Ui Ωk mli with i = 1, 2, 3, 4, 5, 6


1. 2.
and j, k, l = 1, 2, 3

• Moments

Mj = −U̇i mj+3,i − Ejkl Ui Ωk ml+3,i − Ejkl Uk Ui mli with i = 1, 2, 3, 4, 5, 6


3. 2. 3.
and j, k, l = 1, 2, 3

7
Einstein’s Σ notation applies.



⎪ 0 if any j, k, l are equal


⎨ 1 ifj, k, l are in cyclic order, i.e.,
Ejkl = ‘alternating tensor’ = (1, 2, 3), (2, 3, 1), or (3, 1, 2)



⎪ −1 ifj, k, l are not in cyclic order i.e.,

(1, 3, 2), (2, 1, 3), (3, 2, 1)

Note:

(a) if Ωk ≡ 0 , Fj = −U̇i mji (as expected by definition of mij ).

Also if U̇i ≡ 0, then Fj = 0 for any Ui , no force in steady translation.

(b) Bl ∼ Ui mli ‘added momentum’ due to rotation of axes.


  
Then all the terms marked as 2. are proportional to ∼ Ω × B where B is linear
momentum (momentum from i coordinate into new xj direction).

(c) If Ωk ≡ 0 : Mj = −U̇i mj+3,i mij − Ejkl Uk Ui mli .


 
 

even with U˙ =0, Mj =0 due to this term

Moment on a body due to pure steady translation – ‘Munk’ moment.

3.19.5 Example Generalized motions, forces and moments.

A certain body has non-zero added mass coefficients only on the diagonal, i.e. mij =
δij . For a body motion given by U1 = t, U2 = −t, and all other Ui , Ωi = 0, the
forces and moments on the body in terms of mi are:
F1 = , F2 = , F3 = , M1 = , M2 = , M3 =

Solution:

mij = δij

U1 = t U2 = −t Ui = 0 i = 3, 4, 5, 6 Ωk = 0 k = 1, 2, 3

U̇1 = 1 U̇2 = −1 U̇i = 0 i = 3, 4, 5, 6

Use the relations from (JNN §4.13):


k Ω =0
Fj = −U̇i mij − Ejkl Ui Ωk mil −→
Fj = −U̇i mij

k Ω =0
Mj = −U̇i mi(j+3) − Ejkl Ui Ωk mi(l+3) − Ejkl Uk Ui mli −→
Mj = −U̇i mi(j+3) − Ejkl Uk Ui mli

where i = 1, 2, 3, 4, 5, 6 and j, k, l = 1, 2, 3

For F1 , F2 , F3 use the previous relationship for Fj with j = 1, 2, 3 respectively:

F1 = − U̇1 m11 − U̇2 m21 −U̇3 m31 −U̇4 m41 −U̇5 m51 −U̇6 m61 → F1 = −m11
     
=1 =0 =0 =0 =0 =0
Check
F2 = − U̇2 m22 → F2 = m22

=−1

Check
F3 = − U˙ 3 m33 → F3 = 0

=0

9
For M 1, M2 , M3 use the previous relationship for Mj with j = 1, 2, 3 respectively:

M1 = −U̇i mi(1+3) − E1kl Uk Ui mli


= −U̇i mi4 − E1kl Uk Ui mli

= −U̇1 m14 −U̇2 m24 −U̇3 m34 − U̇4 m44 − U̇5 m54 −U̇6 m64

     


=0 =0 =0 =0 =0 =0
 
−E123 U2 U1 m13 +U2 m23 + U3 m33 + U4 m43 +U5 m53 +U6 m63
     
=0 =0 =0 =0 =0 =0
 
−E132 U3 U1 m12 + U2 m22 + U3 m32 +U4 m42 +U5 m52 +U6 m62 → M1 = 0
      
=0 =0 =−1 =0 =0 =0 =0

M2 = −U̇i mi5 − E2kl Uk Ui mli


= U̇5 m55 − E231 U3 Ui m1i − E213 U1 Ui m3i
= −E213 U1 U3 m33 → M2 = 0

M3 = −U̇i mi6 − E3kl Uk Ui mli


= U̇6 m66 − E312 U1 Ui m2i − E321 U2 Ui m1i
 
+1 −1

= − U1 U2 m22 + U2 U1 m11 → M3 = t2 (m22 − m11 )


   
t −t −t t

10

3.19.6 Example Munk Moment on a 2D submarine in steady translation

2
1

θ U

3 (out of page)

U1 =U cos θ
U2 = − U sin θ

Consider steady translation motion: U̇ = 0; Ωk = 0. Then

M3 = −E3kl Uk Ui mli

For a 2D body, m3i = mi3 = 0, also U3 = 0, i, k, l = 1, 2. This implies that:

M3 = − E312 U1 (U1 m21 + U2 m22 ) − E321 U2 (U1 m11 + U2 m12 )


 
=1 =−1

= −U1 U2 (m22 − m11 )


⎛ ⎞

= U 2 sin θ cos θ ⎝m22 − m11 ⎠


  
>0

Therefore, M3 > 0 for 0 < θ < π/2 (‘Bow up’). Therefore, a submarine under
forward motion is unstable in pitch (yaw). For example, a small bow-up tends to
grow with time, and control surfaces are needed as shown in the following figure.

11

B
H

• Restoring moment ≈ (ρg∀H)sinθ.


• critical speed Ucr given by:

2
(ρg∀) H sin θ ≥ Ucr sin θ cos θ (m22 − m11 )

θ
Ucr
H

2
Usually m22 >> m11 , m22 ≈ ρ∀. For small θ, cos θ ≈ 1. So, Ucr ≤ gH or Fcr ≡
U
√ cr
gH
≤ 1. Otherwise, control fins are required.

12

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 14

13.021 - Marine Hydrodynamics


Lecture 14

3.20 Some Properties of Added-Mass Coefficients


1. mij = ρ·[function of geometry only]

F, M = [linear function of mij ] × [function of instantaneous U, U˙ , Ω]


not of motion history

2. Relationship to fluid momentum.

F(t)

where we define Φ to denote the velocity potential that corresponds to unit velocity
U = 1. In this case the velocity potential φ for an arbitrary velocity U is φ = U Φ.
The linear momentum L  in the fluid is given by
   
 =
L ρvdV = ρ∇φdV = + ρφˆ
ndS

V V Green’s
theorem
B ∞

φ→0 at ∞

 
Lx (t = T ) = ρU Φnx dS = U ρΦnx dS
B B

The force exerted on the fluid from the body is −F (t) = −(−mA U̇ ) = mA U̇ .

1
T T Newton’s Law 

dt [−F (t)] = mA U˙ dt = mA U ]T0 = Lx (t = T )−Lx (t = 0) = U ρΦnx dS
  
0 0 mA U B

Therefore, mA = total fluid momentum for a body moving at U = 1 (regardless of


how we get there from rest) = fluid momentum per unit velocity of body.

∂φ ∂φ ∂U Φ ∂Φ
K.B.C. ∂n
= ∇φ · n̂ = (U, 0, 0) · n̂ = U nx , ∂n
= U nx ⇒ ∂n
= U nx ⇒ = nx
∂n

∂Φ
∴ mA = ρ Φ dS
∂n
B

For general 6 DOF:

 
∂Φj
mji =ρ Φi nj dS = ρ Φi dS = j fluid momentum due to
  ∂n
j−force/moment B potential due to body B i body motion
i−direction of motion moving with Ui =1

3. Symmetry of added mass matrix mij = mji .

    
∂Φj
mji = ρ Φi dS = ρ Φi (∇Φj · n̂)dS = ρ ∇ · (Φi ∇Φj ) dV
∂n ↑
Divergence
B B V
⎛ ⎞ Theorem


=ρ ⎝∇Φi · ∇Φj + Φi ∇2 Φj ⎠dV
  
V =0

Therefore,

mji = ρ ∇Φi · ∇Φj dV = mij
V

2
4. Relationship to the kinetic energy of the fluid. For a general 6 DoF body motion
Ui = (U1 , U2 , . . . , U6 ),

φ = Ui Φi ; Φi = potential for Ui = 1

notation


1
K.E. = ρ ∇φ · ∇φdV = 12 ρ Ui ∇Φi · Uj ∇Φj dV
2 V
V

1
= ρUi Uj ∇Φi · ∇Φj dV = 12 mij Ui Uj
2
V

K.E. depends only on mij and instantaneous Ui .

5. Symmetry simplifies mij . From 36 → 21 → ‘?’. Choose such coordinate system


symmetry
that some mij = 0 by symmetry.

Example 1 Port-starboard symmetry.

⎡ ⎤
m11 m12 0 0 0 m16 Fx
⎢ m22 0 0 0
m26 ⎥ Fy
⎢ ⎥
⎢ m33 m34 m35 0 ⎥ Fz
mij = ⎢




m44 m45 0

Mx 12 independent coefficients

m55 0

My
m66 Mz
U1 U2 U3 Ω1 Ω2 Ω3

3
Example 2 Rotational or axi-symmetry with respect to x1 axis.

⎡ ⎤
m11 0 0 0 0 0
⎢ m22 0 0 0 m35 ⎥
⎢ ⎥
⎢ m22 0 m35 0 ⎥ where m22 = m33 , m55 = m66
mij = ⎢


⎥ and m26 = m35 , so 4 different coefficients

0 0 0


m55 0 ⎦

m55

Exercise How about 3 planes of symmetry (e.g. a cuboid); a cube; a sphere?? Work
out the details.

3.21 Slender Body Approximation


Definitions
(a) Slender Body = a body whose characteristic length in the longitudinal direction is
considerably larger than the body’s characteristic length in the other two directions.
(b) mij = the 3D added mass coefficient in the ith direction due to a unit acceleration in
the j th direction. The subscripts i, j run from 1 to 6.
(c) Mkl = the 2D added mass coefficient in the k th direction due to a unit acceleration in
the lth direction. The subscripts k, l take the values 2,3 and 4.

x2
x5
x2
L

x4
x1
O
x 4 x3
x
x6
x3

Goal To estimate the added mass coefficients mij for a 3D slender body.

Idea Estimate mij of a slender 3D body using the 2D sectional added mass coefficients
(strip-wise Mkl ). In particular, for simple shapes like long cylinders, we will use known 2D
coefficients to find unknown 3D coefficients.

mij = [Mkl (x) contributions]
3D 2D

Discussion If the 1-axis is the longitudinal axis of the slender body, then the 3D added
mass coefficients mij are calculated by summing the added mass coefficients of all the thin
slices which are perpendicular to the 1-axis, Mkl . This means that forces in 1-direction
cannot be obtained by slender body theory.

5
Procedure In order to calculate the 3D added mass coefficients mij we need to:

1. Determine the 2D acceleration of each crossection for a unit acceleration in the ith
direction,
2. Multiply the 2D acceleration by the appropriate 2D added mass coefficient to get the
force on that section in the j th direction, and
3. Integrate these forces over the length of the body.

Examples

• Sway force due to sway acceleration


Assume a unit sway acceleration u̇3 = 1 and all other uj , u̇j = 0, with j = 1, 2, 4, 5, 6.
It then follows from the expressions for the generalized forces and moments (Lecture
12, JNN §4.13) that the sway force on the body is given by

f3 = −m33 u̇3 = −m33 ⇔ m33 = −f3 = − F3 (x)dx
L

A unit 3 acceleration in 3D results to a unit acceleration in the 3 direction of each


2D ‘slice’ (U̇3 = u̇3 = 1). The hydrodynamic force on each slice is then given by
F3 (x) = −M33 (x)U̇3 = −M33 (x)

Putting everything together, we obtain


 
m33 = − −M33 (x)dx = M33 (x)dx
L L

• Sway force due to yaw acceleration


Assume a unit yaw acceleration u̇5 = 1 and all other uj , u̇j = 0, with j = 1, 2, 3, 4, 6.
It then follows from the expressions for the generalized forces and moments that the
sway force on the body is given by

f3 = −m35 u̇5 = −m35 ⇔ m35 = −f3 = − F3 (x)dx
L

For each 2D ‘slice’, a distance x from the origin, a unit 5 acceleration in 3D, results
to a unit acceleration in the -3 direction times the moment arm x (U̇3 = −xu̇5 = −x).
The hydrodynamic force on each slice is then given by
F3 (x) = −M33 (x)U̇3 = xM33 (x)

6
Putting everything together, we obtain

m35 = − xM33 (x)dx
L

• Yaw moment due to yaw acceleration


Assume a unit yaw acceleration u̇5 = 1 and all other uj , u̇j = 0, with j = 1, 2, 3, 4, 6.
It then follows from the expressions for the generalized forces and moments that the
yaw force on the body is given by

f5 = −m55 u̇5 = −m55 ⇔ m55 = −f5 = − F5 (x)dx
L

For each 2D ‘slice’, a distance x from the origin, a unit 5 acceleration in 3D, results
to a unit acceleration in the -3 direction times the moment arm x (U̇3 = −xu̇5 = −x).
The hydrodynamic force on each slice is then given by

F3 (x) = −M33 (x)U̇3 = xM33 (x)

However, each force F3 (x) produces a negative moment at the origin about the 5 axis

F5 (x) = −xF3 (x)

Putting everything together, we obtain



m55 = x2 M33 (x)dx
L

In the same manner we can estimate the remaining added mass coefficients mij - noting
that added mass coefficients related to the 1-axis cannot be obtained by slender
body theory.

In summary, the 3D added mass coefficients are shown in the following table. The empty
boxes may be filled in by symmetry.

⎡ ⎤

⎢ m22 = M22 dx m23 = M23 dx m24 = M24 dx m25 = −xM23 dx m26 = xM22 dx ⎥
⎢ ⎥
⎢ L L L L L ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ m33 = M33 dx m34 = M34 dx m35 = −xM33 dx m36 = xM32 dx ⎥
⎢ L L L L ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ m44 = M44 dx m45 = −xM34 dx m46 = xM24 dx ⎥
⎢ ⎥
⎢ L L L ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ m55 = x2 M33 dx m56 = −x2 M32 dx ⎥
⎢ ⎥
⎢ L L ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥
⎢ m66 = x2 M22 dx ⎥
⎣ L

3.22 Buoyancy Effects Due to Accelerating Flow


Example Force on a stationary sphere in a fluid that is accelerated against it.

⎛ ⎞
⎜ a3 ⎟
φ (r, θ, t) = U (t) ⎜
⎝r +

⎠ cos θ
2r2

dipole for sphere

∂φ  3a
 = U˙ cos θ
∂t r=a 2
 
3
∇φ|r=a = 0, − U sin θ, 0
2
1  9
|∇φ|2 r=a = U 2 sin2 θ
2 8
Then,

π  
  3a 9
Fx = (−ρ) 2πr2 dθ sin θ (− cos θ) U̇ cos θ + U 2 sin2 θ
2 8
0
π π
3 2 2 9π 2
= U̇ 3πρa dθ sin θ cos θ +ρU a dθ cos θ sin3 θ
4
0   0  
=2/3 =0
 
Fx = U˙ ρ 2πa3
  
4
3
πa3 ρ + 23 πa3 ρ

=ρ∀

9
Part of Fx is due to the pressure gradient which must be present to cause the fluid to
accelerate:

∂U ∂U ∂U ∂U 1 ∂p
x-momentum, noting U = U (t) : +U +v +w =− (ignore gravity)
∂t ∂x
 ∂y   ∂z ρ ∂x
  
0 0 0
dp
= −ρU̇ for uniform (1D) accelerated flow
dx
Force on the body due to the pressure field
  
 ∂p
F = pndS
ˆ =− ∇pdV ; Fx = − dV = ρ∀U̇
∂x
B VB VB

‘Buoyancy’ force due to pressure gradient = ρ∀U̇

10

Analogue: Buoyancy force due to hydrostatic pressure gradient. Gravitational accelera­


tion g ↔ U̇ = fluid acceleration.

ps = −ρgy

∇ps = −ρg ĵ → F s = −ρg∀ĵ Archimedes principle
Summary: Total force on a fixed sphere in an accelerated flow
 
Fx = U̇ ρ∀ + m(1) = U˙ 32 ρ∀ = 3U˙ m(1)
 
Buoyancy added mass
1
2
ρ∀

In general, for any body in an accelerated flow:

Fx = Fbuoyany + U˙ m(1) ,
where m(1) is the added mass in still water (from now on, m)

Fx = −U˙ m for body acceleration with U̇

Added mass coefficient


m
cm =
ρ∀
in the presence of accelerated flow Cm = 1 + cm

11

Appendix A: More examples on symmetry of added mass tensor

•Symmetry with respect to Y (= “X-Z” plane symmetry) 12 non-zero, independent coefficients

•Symmetry with respect to X and Y (= “Y-Z” and “X-Z” plane symmetry) 7 non-zero, independent coefficients

12
•Axisymmetric with respect to X-axis 4 non-zero, independent coefficients

•Axisymmetric with respect to X axis and X (=“Y-Z” plane symmetry) 3 non-zero, independent coefficients

13
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 15

13.021 - Marine Hydrodynamics


Lecture 15

 0)
Chapter 4 - Real Fluid Effects (ν =
Potential Flow Theory → Drag = 0.

Observed experiment (real fluid ν << 1 but = 0) → Drag = 0.

In particular the total drag measured on a body is regarded as the sum of two components:
the pressure or form drag, and the skin friction or viscous drag.

Total Drag = Pressure Drag + Skin Friction Drag


Profile Drag or Form Drag or Viscous Drag
Drag Force due to Pressure Drag Force due to Viscous Stresses
� �� �� � � �� �� �
pn̂ds τ t̂ds
S S

where n̂ and t̂ are the normal and tangential unit vectors on the body surface respectively.
The pressure and the viscous stresses on the body surface are p and τ respectively.

The form drag is evaluated by integrating the pressure along the surface of the body. For
bluff bodies that create large wakes the form drag is ∼ total drag.

The skin friction drag is evaluated by integrating the viscous stresses on and along the
body boundary. For streamlined bodies that do not create appreciable wakes, friction drag
is dominant.

4.1 Form Drag


4.1.1 Form Drag on a Bluff Body
Consider a sphere of diameter d:
D (Drag)
ν
U
d
ρ

If no DBC apply then we have seen from Dimensional Analysis that the drag coefficient
is a function of the Reynolds number only:

CD = CD (Re )
The drag coefficient CD is defined with respect to the body’s projected area S:
D D
CD = 1 =
ρU 2 S 1
ρU 2 πd2 /4
2 2 � �� �
Projected area

The Reynolds number Re is defined with respect to the body’s diameter d:


Ud
Re =
ν
The following graph shows the dependence of CD on Re as measured from numerous ex­
periments on spheres.

CD

0.5

0.25

Re
3x105

Figure 1: Drag coefficient (CD ) for a sphere for Re > 102 .

i) Bluff Body → Form Drag


For a bluff body (examples: sphere, cylinder, flat plate, etc.) there is appreciable flow
separation and a wake is formed downstream of the body. The pressure within the
wake is significantly smaller than that upstream of the body. Therefore the integral
of the pressure along the body boundary (= form drag) does not evaluate to zero as
predicted by P-Flow.

In general, for bluff bodies form drag >> friction drag

Flow separation

Real Flow Potential Flow

Assume: pressure in the wake ∼ p∞ (pressure at infinity)


pressure on the upstream boundary of the body ∼ ps (stagnation pressure)
Then:
� � � 1 2�
D= CD (Projected/frontal area)(ps − p∞ ) = CD S ρU
���� � �� � ↑ 2
Friction Coefficient Bernoulli
S
to be determined

ii) CD = CD (Re ) → Regime Dependence


In general, for typical bluff bodies such as spheres, it is found that CD = CD (Re )
has a form similar to that shown in Figure (1). This means that CD has a ‘regime
dependence’ on Re .

There are two main regimes of interest:

• Laminar regime: Re no separation < Re < Re critical ∼


= ( 3 × 105 )
for a smooth sphere with smooth inflow

Laminar boundary layer


Separation pt

• Wide wake
No Stagnation pt Wake • Early separation
Drag
Stagnation pt
Width ~ Diameter • ‘Large’ CD =O(1)

Separation pt
Laminar boundary layer

• Turbulent regime: Re > Re critical

Turbulent boundary layer


Separation pt

No Stagnation pt Wake • Narrow wake


Width ~ Diameter/2 • Delayed separation
Stagnation pt
Separation pt
• Smaller CD

Turbulent boundary layer

D/L
iii) Cylinder The drag coefficient for a cylinder is defined as:
CD = 1
2
ρU 2 d

D
U

L d

CD

1.2

0.6

Re
3x105

Figure 2: Drag coefficient (CD ) for a cylinder for Re > 102

iv) Bodies with Fixed Separation Points


For bodies with fixed separation points, the drag coefficient is roughly constant, i.e.,
does not depend on Re . For example, for a flat plate or disc CD ≈ 1.2

Separation pt

Separation pt

4.1.2 Boundary Layers


� �∗
L

∂v � �
 ∗ ν � 2 �∗
+ v · ∇v = ... + ∇ v
UT ∂t UL
����
1
Re
L

For most flows of interest to us ReL >> 1, i.e., viscosity can be ignored if U, L govern
the problem, thus potential flow can be assumed. In the context of potential flow theory,
drag = 0! Potential flow (no τij ) allows slip at boundary, but in reality, the no-slip condi­
tion applies on the boundaries. Otherwise, if ν =  0 and a free-slip KBC is imposed then
∂u
τ ∼ ν ∂y → ∞ at the boundary.

Prandtl: There is a length scale δ (boundary layer thickness δ << L) over which
velocity goes from zero on the wall to the potential flow velocity U outside the boundary
layer.

U
u=U

U δ<<L
y
L x

u=0

Estimate δ: Inside the boundary layer, viscous effects are of the same order as the
inertial effects.


∂ 2U ∂U U U2 ν δ2 δ ν 1
ν 2 ∼U →ν 2 ∼ → ∼ 2 → ∼ =� << 1 As ReL ↑, δ ↓
∂y ∂x δ L UL L L UL ReL

Generally: ReL >> 1, Lδ << 1, thus potential flow is good outside a very thin bound­
ary layer (i.e., provided no separation - a real fluid effect). For Reynolds number not
>> 1(Re ∼ O(1)), then thick boundary layer ( δ ∼ O(L)) and Prandtl’s boundary layer
idea not useful. If separation occurs, then boundary layer idea is not valid.

6
4.1.3 Boundary Layers and Flow Separation

Boundary layers help understand flow separation.

Example for flow past a circle.

Outside the boundary layer P-Flow is valid. Let capital U denote the potential flow

tangential velocity on the circle and let x denote the distance along the circle surface (i.e.,

x = body coordinate).

From the steady inviscid x-momentum equation (steady Euler) along the body boundary
(y = 0, V = 0), we obtain :
dU 1 dp
U =− (1)
dx ρ dx

Note 1: Equation (1) is used frequently in boundary layer theory.

dp
Note 2: From Equation (1) → for flow past a flat plate dx = 0 along the plate.

dU 1 dp
P − Flow solution on body y=0 :U =−
dx dx

y U = U max boundary layer : u, υ

x δ
U0
ν

U =0 U =0

dU dU
>0 <0
dx dx
dp dp
<0 >0
dx dx

dU dU
dx
>0 Acceleration dx
<0 Deceleration

dp dp
dx
<0 ‘Favorable’ pressure gradient dx
>0 ‘Adverse’ pressure gradient

X2 X3
X4
X1
X5

X2 > X1
X=X1 y
y

P
P
p
p

u
u v
v U1 ω U2 > U1
ω
P>p Flow is being
pushed to attach

X3 > X2 X4 > X3
y y

P
P
p

u u
U3 U2 ∂u v U4 U3
= 0, ω = 0
τ3 > 0 ∂y τ4 = 0
X4 is defined as the point of
separation

X5 > X4
y

u
U5 U4

Separated Flow τ5 < 0

Flow reversal

A better way to think about separation is in terms of diffusion of vorticity.

ω=0 outside B.L.


y y

P P

ω2 ω(y) ω4 = 0 ω(y)
ω1 ω3
ω removed from fluid
ω added to fluid
by diffusion

Think of vorticity as heat; ω(y) is equivalent to a temperature distribution. Note:




DV  Dω
= ... + ν∇2 V and = ... + ν∇2 ω


Dt Dt

13.021 – Marine Hydrodynamics, Fall 2004


Lecture 16

13.021 - Marine Hydrodynamics


Lecture 16

4.1.4 Vortex Shedding and Vortex Induced Vibrations


Consider a steady flow Uo over a bluff body with diameter D.

Fy (lift)

U0
D Fx (drag)

We would expect the average forces to be: However, the measured oscillatory forces are:

F
F
Fx
Average

Fx
Average

Fy t
Fy
t

• The measured drag Fx is found to oscillate about a non-zero mean value with
frequency 2f .

• The measured lift Fy is found to oscillate about a zero mean value with frequency
f.

• f = ω/2π is the frequency of vortex shedding or Strouhal frequency.

1
Reason: Flow separation leads to vortex shedding. The vortices are shed in a staggered
array, within an unsteady non-symmetric wake called von Karman Street. The frequency
of vortex shedding is the Strouhal frequency and is a function of Uo , D, and ν.

Fy Von
Karman
Street
Uo Fx
D

Strouhal frequency
����
f D
i) Strouhal Number We define the (dimensionless) Strouhal number S ≡ .
U0

The Strouhal number S has a regime dependence on the Re number S = S(Re ).

For a cylinder:


Laminar flow S ∼ 0.22

S(Re)

Turbulent flow S ∼ 0.3

0.3
0.22

105 106 107 Re

ii) Drag and Lift The drag and lift coefficients CD and CL are functions of the correlation
length.
For ‘∞’ correlation length:

• If the cylinder is fixed, CL ∼ O(1) comparable to CD .


• If the cylinder is free to move, as the Strouhal frequency fS approaches one of
the cylinder’s natural frequencies fn , ‘lock-in’ occurs. Therefore, if one natural
frequency is close to the Strouhal Frequency fn ∼ fS , we have large amplitude
motions ⇒Vortex Induced Vibration (VIV).

2
4.2 Drag on a Very Streamlined Body

UL
R eL ≡
ν Uo D
D b
Cf ≡ 1
ρU 2 (Lb)
2 ���� L
S=wetted area
one side of plate

Cf = Cf (ReL , L/b)
� �
Unlike a bluff body, Cf is a strong function of ReL since D is proportional to ν τ = ν ∂u
∂y
.
See an example of Cf versus ReL for a flat plate in the figure below.

Cf

Laminar
0.01
Turbulent

0.001

Re
105 106

Skin friction coefficient as a function of the Re for a flat plate

• ReL depends on plate smoothness, ambient turbulence, . . .

• In general, Cf ’s are much smaller than CD ’s (Cf /CD ∼ O(0.1) to O(0.01)). Therefore,
designing streamlined bodies allows minimal separation and smaller form drag at the
expense of friction drag.

• In general, for streamlined bodies CTotal Drag is a combination of CD (Re ) and Cf (ReL ),
� �
1 2
and the total drag is D = 2 ρU CD S + C f Aw , where CD has a regime
frontal area wetted area
dependence on Re and Cf is a continuous function ReL .

3
4.3 Known Solutions of the Navier-Stokes Equations
4.3.1 Boundary Value Problem

• Navier-Stokes’:

∂�v 1 1
+ (�v · �) �v = − �p + ν�2�v + f�
∂t ρ ρ

Conservation of mass:

� · �v = 0

• Boundary conditions on solid boundaries “no-slip”:


�v = U

Equations very difficult to solve, analytic solution only for a few very special cases (usually
when �v · ��v = 0. . . )

4.3.2 Steady Laminar Flow Between 2 Long Parallel Plates: Plane Couette Flow

Steady, viscous, incompressible flow between two infinite plates. The flow is driven by a
pressure gradient in x and/or motion of the upper plate with velocity U parallel to the
x-axis. Neglect gravity.

Assumptions Governing Equations Boundary Conditions

∂ ∂u ∂v ∂w
i. Steady Flow: ∂t =0 Continuity: ∂x + ∂y + ∂z =0 �v = (0, 0, 0) on y = 0
∂�v ∂�v ∂�v
ii. (x, z) >> h: ∂x = ∂z =0 NS: ∂t + �v · ��v = − ρ1 �p + ν�2�v �v = (0, 0, 0) on y = h

iii. Pressure: independent of z

Continuity
∂u ∂v ∂w ∂v
+ + =0⇒ = 0 ⇒ v = v(x, z) ⇒ v = 0 (1)
∂x ∂y ����
���� ∂z ∂y ↑
BC: v(x,0,z)=0
=0, from assumption ii

Momentum x
� 2 �
∂u ∂u ∂u ∂u 1 ∂p ∂ u ∂ 2u ∂ 2u
+u + ����
v +w =− +ν + + 2 ⇒
∂t
���� ∂x
���� ∂y ∂z
���� ρ ∂x ∂x2 ∂y 2 ����
���� ∂z
=0, (1)
=0, i =0, ii =0, ii =0, ii =0, ii
2
∂ u 1 ∂p
ν 2
= (2)
∂y ρ ∂x

Momentum y
∂v 1 ∂p
+�v · � ����
v =− + ν�2 ����
v

∂t
���� ρ ∂y
=0, (1) =0, (1)
=0, i

∂p ∂p dp
=0 ⇒ p = p(x) and = (3)
∂y ↑ ∂x dx
assumption iii

Momentum z
� 2 �
∂w ∂w ∂u ∂w 1 ∂p ∂ w ∂ 2w ∂ 2w
+u + ����
v +w =− +ν + + ⇒
∂t
���� ∂x
���� ∂y ∂z
���� ρ ����
∂z ∂x2 ∂y 2 �∂z
���� ��
2

=0, (1)
=0, i =0, ii =0, ii =0, iii =0, ii =0, ii
2
∂ w
= 0 ⇒ w = ay + b ⇒ w=0 (4)
∂y 2 ↑
w(x,0,z)=0
w(x,h,z)=0

From Equations (1), (4)


∂u du
�v = (u, 0, 0). Also ⇒ u = u(y) and = (5)
↑ ∂y dy
assumption ii

From Equations (2), (3, and (5)


� � � �
d2 u 1 d2 p 1 dp 2 1 dp y
= ⇒ u=− − y +C1 y+C2 ⇒ u= − (h − y)y + U
dy 2 ρν dx2 ↑ 2µ dx ↑ 2µ dx h
µ=ρν u(x,0,z)=0

w(x,h,z)=U

• Special cases for Couette flow


dp Px −Px+L
1
u(y) = 2µ (h − y)y(− dx ) + U hy , where (− dx
dp
)= L

� dp
� � dp

I. U = 0, − dx
>0 II. U �= 0, − dx
=0

y y

dp dp
(− )>0 (− )=0
dx u ( h) = U = 0 dx u ( h) = U U

h p p
h

u ( y) u ( y)
u ( 0) = 0 u ( 0) = 0

Parabolic profile Linear profile

•Velocity
dp
u(y) = 1

(h − y)y(− dx ) u(y) = U hy
•Max velocity
h2 dp
umax = u(h/2) = 8µ
(− dx ) umax = U
•Volume flow rate
�h h3 dp
Q = 0 u(y)dy = 8µ (− dx ) Q = h2 U
•Average velocity
Q h2 dp U
ū = h
= 6µ
(− dx ) ū = 2

•Viscous stress on bottom plate (skin friction)


� � dp � �
du � du �
τw = µ dy � = h2 − dx >0 τw = µ dy �
= µ Uh
y=0 y=0

� dp

III. U �= 0, − dx
�= 0

u ( h) = U U u ( h) = U U
back flow
h
τw τw

dp dp
U > 0, (− ) > 0, G > 0 U > 0, (− ) < 0, G < 0
dx dx

� dp
� � dp

− dx
>0 − dx
<0
•Viscous stress on bottom plate (skin friction)
� dp �
τw = h2 − dx + µ Uh

< dp < 2µU attached
τw =0 when (− )= − 2 , in which case the flow is insipient
> dx > h separated

� dp

For the general case of U �= 0 and − dx
�= 0,
h � dp � U
− +µ τw =
2 dx h
We define a Dimensionless Pressure Gradient G

h2 � dp �
G≡ −
2µU dx
such that

• G > 0 denotes a favorable pressure gradient


• G < 0 denotes an adverse pressure gradient
• G = −1 denotes an incipient flow
• G < −1 denotes a separated or back-flow

u ( h) = U U u ( h) = U U u ( h) = U U

h
back flow

dp dp dp
U > 0, (− ) > 0, G > 0 U > 0, (− ) < 0, G < 0 U > 0, (− ) < 0, G = −1 ⇒ τ w = 0
dx dx dx

Lessons learned in § 4.3.2:

1. Reviewed how to simplify the Navier-Stokes equations.

2. Obtained one solution to the Navier-Stokes equations.

3. Realized that once the Navier-Stokes are solved we know


everything.

In the next paragraph we are going to study one more solution to the Navier-Stokes equa­
tion, in polar coordinates.

9
4.3.3 Steady Laminar Flow in a Pipe: Poiseuille Flow
∂x
y

a
x

z
r=a
L
Vx(r)
Steady, laminar pipe flow. KBC: vx (a) = 0 (no slip) and
dvx
(r2 = y 2 + z 2 , �v = (vx , vr , vθ )) dr (0) = 0 (symmetry).

Assumptions Governing Equations Boundary Conditions

∂ 1 ∂rvr 1 ∂vθ ∂vx


i. Steady Flow: ∂t =0 Continuity: r ∂r + r ∂θ + ∂x =0 vx (r = a) = 0 no-slip
∂�v ∂�v dvx
ii. (x, z) >> h: ∂x = ∂θ =0 NS: In polar coordinates (see SAH pp.74) dr |r=0 = 0 symmetry

⇒ �v = �v (r)

iii. Pressure: independent of θ

Following a procedure similar to that for plane Couette flow (left as an exercise) we can
show that
� � ��
1 dp 1 d dvx
vr = vθ = 0, vx = vx (r), p = p(x), and =ν r
ρ dx r dr dr
� �� �
r component of �2
in cylindrical coordinates

After applying the boundary conditions we find:


� �
1 dp � 2 �
vx (r) = − a − r2
4µ dx
Therefore the volume flow rate is given by

� 2π � a
π 4 � dp �
Q= dθ rdrvx (r) = a −
0 0 8µ dx
and the skin friction evaluates to
� �
∂vr ∂vx �� ∂vx �� a � dp �
τw = τx (−r) = −τxy = −µ ( + )� = −µ � ⇒ τw = −
∂x ∂r r=a ∂r r=a 2 dx

10

4.4 Boundary Layer Growth Over an Infinite Flat Plate for


Unsteady Flow
Boundary layer thickness is related to the area where the viscosity and vorticity effects are
diffused.

For a flow over an infinite flat plate, the boundary layer thickness increases unless it is
constrained in the y direction and/or by time (unsteady flow).

1. Steady flow, constrained in y


For a steady flow past a flat plate, the boundary layer thickness increases with x.
If the flow is constrained in y, eventually the viscous effects are diffused along the
entire cross section and the flow becomes invariant in the streamwise direction.
In paragraphs 4.3.2 and 4.3.3, we studied two cases of steady laminar viscous flows,
where the viscous effects had diffused along the entire cross section.

y
h
U0
x
boundary layer Couette flow for x >> h
thickness increases with x

� � � �
Couette h
Steady flow, we assumed that viscous effects diffused through entire .
Poiseuille a

11

2. Unsteady flow, unconstrained in y

Consider the simplest example of an infinite plate in unsteady motion.

U(t)
x

∂�v ∂�v
Assumptions �p = 0, ∂x
= ∂z
= 0 ⇒ �v = �v (y, t)

Can show v = w = 0 and u = u(y, t).


Finally, from u momentum (Navier-Stokes in x) we obtain

⎛ ⎞
2 2 2
∂u ∂u ∂u ∂u 1 ∂p ⎜∂ u ∂ u ∂ u⎟
+u + ����
v +w =− +ν ⎝ 2 + 2 + 2 ⎠ ⇒
∂t ∂x
���� ∂y � ��
∂z� ρ ����
∂x ∂x
���� ∂y ∂z
����
=0
=0 =0 =0 =0 =0

∂u ∂ 2u
=ν 2 ‘ �momentum
�� � ’ diffusion equation (6)
∂t ∂y velocity
(heat)

Equation (6) is:

� first order PDE in time → requires 1 Initial Condition


� second order PDE in y → requires 2 Boundary Conditions
- u(y, t) = U (t) at y = 0, for t > 0
- u(y, t) →
0 as y → ∞

From Equation (6), we observe that the flow over a moving flat plate is due to viscous
dissipation only.

12
4.5.1 Sinusoidally Oscillating Plate

i. Evaluation of the Velocity Profile for Stokes Boundary Layer

The flow over an oscillating flat plate is referred to as ‘Stokes Boundary Layer’.

Recall that eiα = cos α + i sin α where α is real.

Assume that the plate is oscillating with U (t) = Uo cos ωt = Real {Uo eiωt }. From linear
theory, it is known that the fluid velocity must have the form
� �
u (y, t) = Real f (y) eiωt , (7)

where f(y) is the unknown complex (magnitude & phase) amplitude of oscillation.

To obtain an expression for f (y), simply substitute (7) in (6). This leads to:

d2 f
iωf = ν 2 (8)
dy

Equation (8) is a 2nd order ODE for f (y). The general solution is
√  √ 
(1+i) ω/2ν y −(1+i) ω/2ν y
f (y) = C1 e + C2 e (9)

The velocity profile is obtained from Equations (7), (9) after we apply the Boundary
Conditions.
� √ω �
u(y, t) must be bounded as y → ∞ ⇔ C1 = 0 � ω �
u (y, t) = Uo (e−y 2ν ) cos − y + ωt
u(y = 0, t) = U (t) ⇔ f (y = 0) = Uo ⇔ C2 = Uo 2ν
Stokes Boundary Layer

13

ii. Some Calculations for the Stokes Boundary Layer

Once the velocity profile is evaluated, we know everything about the flow.

ω
−y

e
λ = 2πδ

δ1/ e ≡ δ
u( y)
−1 1/ e 1 Uo
u(y )
Stokes Boundary Layer. Velocity ratio Uo as a function of the distance from the plate y .

Observe: �
u(y, t) √ω �
ω �
−y
= (e 2ν )

cos − y + ωt (10)
Uo � �� � 2ν

Exponentially decaying � �� �
envelope Oscillating component

SBL thickness
The ratio Uuo is composed of an exponentially decaying part → thickness of SBL decays
exponentially with y. We define various parameters that can be used as measures of
the SBL thickness:
u(δ )
• We define δ1/e as the distance y from the plate where U1o/e = 1e . Substituting

into (10), we find that δ ≡ δ1/e = 2νω

• The oscillating component has wave length λ = 2π 2ν = 2πδ. At λ, u(λ) ∼ 0.002.
ω Uo =
u(δ1% )
• We define δ1% as the distance y from the plate, where = 1%. Substituting
� Uo
u(δ1% )
into (10), we find that δ1% = − ln( Uo ) 2ν ∼ 4.6δ.
ω =

14

Numerical examples:
For oscillating plate in water (ν = 10−6 m2 /s= 1mm2 /s) we have
4.6 √ ∼ �
δ1% = √ T = 2.6 ����
T
���� π
in mm in sec


T = ω
δ1%

1s 3mm

10s ≤1cm

Excursion length and SBL


The plate undergoes a motion of amplitude A.
Uo
X = A sin(ωt) ⇒ U = X˙ = ����
Aω cos(ωt) ⇒ ω =
A
Uo

Comparing the SBL thickness ∼ δ with A, we find


� � �
δ ν/ω νA/Uo ν 1
∼ = = ∼�
A A ↑ A Uo A ReA
Uo
ω= A

Skin friction
The skin friction on the plate is given by
� �
∂u �� ω � �
τw = µ � = . . . = µUo sin ωt − cos ωt
∂y y=0 2ν

The maximum skin friction on the wall is



ω
|τw |max = µUo
ν

3π 7π
and occurs at ωt = 4
, 4 ,···

15
4.5.2 Impulsively Started Plate

U(t)

Uo

Recall Equation (6) that describes the the flow u(y, t) over an infinite flat plate undergoing
unsteady motion.
∂u ∂ 2u
=ν 2
∂t ∂y
For an impulsively started plate, the Boundary Conditions are:

u(o, t) = Uo
for t > 0, i.e. u(y, 0) = 0
u(∞, t) = 0

Notice that the problem stated by Equation (6) with the above Boundary Conditions has no
explicit time scale. In this case it is standard procedure to (a) use Dimensional Analysis
to find the similarity parameters of the problem, and (b) look for solution in terms of the
similarity parameters:

u � y � u
u = f (Uo , y, t, ν) ⇒ =f √ ⇒ = f (η) Self similar solution
↑ Uo 2 νt Uo
DA � �� �
≡η
similarity parameter

The velocity profile is thus given by� :


�η
u 2 2
= erfc (η) = 1 − erf (η) = 1 − √ e−α dα
Uo � �� � π
Complementary
0
error function

16


Hints on obtaining the solution:

η = 2√yνt ⎪




⎬ ∂u 2
=ν ∂ u d(u/Uo ) d2 (u/Uo )
∂ ∂η ∂ y ∂ ∂t ∂y 2
∂t
= ∂t ∂η
= − √
4t νt ∂η −→ −η = −→ . . .

⎪ dη dη 2

⎪ � �� �
∂2
� ∂η �2 ∂ 2 ⎪
1 ∂ 2 ⎭ 2nd order ODE
∂y 2
= ∂y ∂η2 = 4νt ∂η2

Boundary layer thickness


In the same manner as for the SBL, we define various parameters that can be used to
measure the boundary layer thickness:

• δ ≡ 2 νt. At y = δ −→ uU(δo) ∼

= 0.16.
• δ1% ∼

= 1.82δ.

Excursion length and boundary layer thickness


At time t, the plate has travelled a distance L = Uo t → t = ULo .
Comparing the boundary layer thickness ∼ δ with L, we find
√ � �

δ νt νL/Uo ν 1


=
=


L L L Uo L ReL

Skin friction
The skin friction on the plate is given by

∂u �
� Uo
τw = µ � = . . . = −µ √
∂y y=0 πνt

17

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 17

13.021 - Marine Hydrodynamics


Lecture 17

4.6 Laminar Boundary Layers

y
U potential flow
u, v viscous flow
δ
x

Uo

4.6.1 Assumptions

• 2D flow: w, ≡ 0 and u (x, y) , v (x, y) , p (x, y) , U (x, y).
∂z

• Steady flow: ≡ 0.
∂t

• For δ << L, use local (body) coordinates x, y, with x tangential to the body
and y normal to the body.

• u ≡ tangential and v ≡ normal to the body, viscous flow velocities (used inside
the boundary layer).

• U, V ≡ potential flow velocities (used outside the boundary layer).

1
4.6.2 Governing Equations

• Continuity

∂u ∂v
+ = 0 (1)
∂x ∂y
• Navier-Stokes:

 2 
∂u ∂u 1 ∂p ∂ u ∂2u
u +v =− +ν + (2)
∂x ∂y ρ ∂x ∂x2 ∂y 2
 2 
∂v ∂v 1 ∂p ∂ v ∂ 2v
u +v =− +ν + (3)
∂x ∂y ρ ∂y ∂x2 ∂y 2
4.6.3 Boundary Conditions

• KBC
Inside the boundary layer:
No-slip u(x, y = 0) = 0
No-flux v(x, y = 0) = 0
Outside the boundary layer the velocity has to match the P-Flow solution.

Let y  ≡ y/δ, y ∗ ≡ y/L, and x∗ ≡ x/L. Outside the boundary layer y  → ∞

but y ∗ → 0. We can write for the tangential and normal velocities


u(x∗ , y  → ∞) = U (x∗ , y ∗ → 0) ⇒ u(x∗ , y  → ∞) = U (x∗ , 0),
and v(x∗ , y  → ∞) = V (x∗ , y ∗ → 0) ⇒ v(x∗ , y  → ∞) = V (x∗ , 0) = 0
No-flux
P-Flow

In short:

u(x, y  → ∞) = U (x, 0)

v(x, y  → ∞) = 0
• DBC
As y  → ∞, the pressure has to match the P-Flow solution. The x-momentum
equation at y ∗ = 0 gives
∂U ∂U 1 dp ∂ 2U dp ∂U
U +V =− + ν ⇒ = −ρU
∂x ↓ ∂y ρ dx ↓ ∂y 2 dx ∂x
0
0

2
4.6.4 Boundary Layer Approximation
Assume that ReL >> 1, then (u, v) is confined to a thin layer of thickness δ (x) << L.
For flows within this boundary layer, the appropriate order-of-magnitude scaling /
normalization is:

Variable Scale Normalization

u U u = Uu∗

x L x = Lx∗

y δ y = δy ∗

v V =? v = Vv ∗

• Non-dimensionalize the continuity, Equation (1), to relate V to U


 ∗  ∗  
U ∂u V ∂v δ
+ = 0 =⇒ V = O U
L ∂x δ ∂y L

• Non-dimensionalize the x-momentum, Equation (2), to compare δ with L


⎡ ⎤
 ∗  ∗ ∗   2 ∗ ⎥
U2 ∂u UV ∂u 1 ∂p νU ⎢⎢ δ2 ∂ 2u ∂ u ⎥
u + v =− + 2 ⎢ 2 + ⎥
L ∂x δ

∂y ρ ∂x δ ⎣ L ∂x 2 ∂y 2 ⎦
2


O(U /L) ignore

The inertial effects are of comparable magnitude to the viscous effects when:

U2 νU δ ν 1
∼ 2 =⇒ ∼ = << 1
L δ L UL R eL
∂p
The pressure gradient ∂x
must be of comparable magnitude to the inertial effects
 2
∂p U
=O ρ
∂x L

3
∂p ∂p
• Non-dimensionalize the y-momentum, Equation (3), to compare ∂y
to ∂x
 ∗  ∗  2 ∗  2 ∗
UV ∂v V2 ∂v 1 ∂p νV ∂ v νV ∂ v
u + v =− + +
L

∂x δ

∂y ρ ∂y L 2

∂x 2 δ 2

∂y 2
2 δ 2 δ 2 δ3 2 δ

O( UL ) O( UL ) O( UL ) O( UL )

L L L3 L

∂p
The pressure gradient ∂y
must be of comparable magnitude to the inertial effects
 2 
∂p U δ
=O ρ
∂y L L
∂p ∂p
Comparing the magnitude of ∂x
to ∂y
we observe
 2   2
∂p U δ ∂p U
= O ρ while =O ρ =⇒
∂y L L ∂x L
∂p ∂p ∂p
<< =⇒ ≈0 =⇒
∂y ∂x ∂y

p = p(x)

• Note:

- From continuity it was shown that V/U ∼ O(δ/L) ⇒ v << u, inside the
boundary layer.
∂p
- It was shown that ∂y = 0, p = p(x) inside the boundary layer. This means that
the pressure across the boundary layer is constant and equal to the pressure
outside the boundary layer imposed by the external P-Flow.

4.6.5 Summary of Dimensional BVP


Governing equations for 2D, steady, laminar boundary layer
∂u ∂v
Continuity : + =0
∂x ∂y
∂u ∂u 1 dp ∂ 2u
x-momentum : u +v = − +ν 2
∂x ∂y ρ dx ∂y


U dU/dx, y=0
∂p
y - momentum : =0
∂y

Boundary Conditions
KBC

At y=0 : u(x, 0) = 0
v(x, 0) = 0
At y/δ → ∞ : u(x, y/δ → ∞) = U (x, 0)
v(x, y/δ → ∞) = 0

DBC
dp ∂U IN the b.l.
1
= −ρU or p(x) = C − ρU 2 (x, 0)
dx ∂x 
2
Bernoulli for the P-Flow at y =0

4.6.6 Definitions
∞ 
∗ u
Displacement thickness δ ≡ 1− dy
U
0

∞
u u
Momentum thickness θ≡ 1− dy
U U
0

Physical Meaning of δ ∗ and θ


Assume a 2D steady flow over a flat plate.

dp
Recall for steady flow over flat plate dx
= 0 and pressure p = const.

Choose a control volume ([0, x] × [0, y/δ → ∞]) as shown in the figure below.

y /δ → ∞ CV
Q
Uo 4
Uo
3
Uo P - Flow

u(y)
Boundary Layer
1
2
0 x
CV for steady flow over a flat plate.

Control Volume ‘book-keeping’

Surface n̂ v v · n̂ v (v · n̂) −pn̂


1 −î Uo î −Uo −Uo2 î pî

2 −ĵ 0 0 0 pĵ

3 î u(x, y)î + v(x, y)ĵ u(x, y) u2 (x, y)î + u(x, y)v(x, y)ĵ −pî

4 ĵ Uo î + v(x, y)ĵ v(x, y) v(x, y)Uo î + v 2 (x, y)ĵ −pĵ

Conservation of mass, for steady CV


  ∞  ∞  x

  
v · ndS
ˆ =0⇒− Uo dy + v(x
, y)dx = 0 ⇒ u(x, y )dy +
1234 0 0
0

Q
 ∞  ∞  ∞  ∞ 
   u
Q= Uo dy − udy = (Uo − u)dy = Uo 1− dy  ⇒ Q = U0 δ ∗
U
0 0 0
0
o

δ

where () are the dummy variables.

Conservation of momentum in x, for steady CV


 
u(v · n
ˆ)dS = Fx ⇒
1234
 ∞  ∞  x  ∞  ∞ 
−Uo2 dy  + 2 
u (x, y )dy + 
v(x , y)Uo dx =  
pdy − pdy  + Fx,f riction ⇒
0 0 0 0 0
 ∞  ∞  x 

−Uo2 dy  + 2  
u (x, y )dy + Uo v(x , y)dx = Fx,f riction ⇒

0 0
0

Q
 ∞  ∞  ∞ 
−Uo2 dy  + u2 (x, y  )dy  + Uo (Uo − u)dy  = Fx,f riction ⇒
0 0 0
 ∞   
− Uo2 +u +2
Uo2 − Uo u dy  = Fx,f riction ⇒
0
 ∞   

u2 u
Uo2 − dy 
= Fx,f riction ⇒

0 Uo2 Uo
  ∞   
2 u u
Fx,f riction = −Uo 1− dy  ⇒ Fx,f riction = −Uo2 θ
Uo Uo
0

θ

4.7 Steady Flow over a Flat Plate: Blasius’ Laminar Boundary


Layer

Uo

L x

Steady flow over a flat plate: BLBL

4.7.1 Derivation of BLBL


dp
• Assumptions Steady, 2D flow. Flow over flat plate → U = U0 , V = 0, =0
dx
• LBL governing equations
∂u ∂v
+ =0
∂x ∂y
∂u ∂u ∂2u
u +v =ν 2
∂x ∂y ∂y
• Boundary conditions
u = v = 0 on y = 0
y 
v → V = 0, u → Uo outside the BL, i.e., >> 1
δ
• Solution Mathematical solution in terms of similarity parameters.
 
u Uo y η νx
and η≡y ⇔ =√ ⇔y=η
U νx x Rx Uo
Similarity solution must have the form

u (x, y)
= F (η)
U
 o

self similar solution

8
We can obtain a PDE for F by substituting into the governing equations. The
PDE has no-known analytical solution. However, Blasius provided a numerical
solution. Once again, once the velocity profile is evaluated we know everything
about the flow.

4.7.2 Summary of BLBL Properties: δ, δ0.99 , δ ∗ , θ, τo , D, Cf


 
u (x, y) Uo νx y η
F (η) ;
=


η = y
;
y ≡ η
;
=


Uo νx
Uo x Rx
evaluated
numerically


local R#



νx ⎪
δ≡ ⎪

Uo ⎪








νx ⎪
⎪ 
δ
.99 ∼
= 4.9 , i.e., η.99 = 4.9 ⎪ ⎪

Uo ⎪

δ ∝ x, δ ∝ 1 Uo

 

∗ ∼ νx
⎪ δ ν
δ = 1.72 , i.e.,
η ∗ = 1.72


⎪ ∝
1

= √

Uo ⎪
⎪ x Uo x Rx







∼ νx ⎪

θ = 0.664 ⎪

Uo ⎪


−1/2 ⎫

U x ⎪

τ o ≡ τw ∼ ⎪
2 o
= 0.332ρUo ⎪

1

ν

τo ∝ √
x



−1/2 ⎪

= 0.332 ρUo2 Rx ⎪
⎪ 3/2



τo ∝ Uo

local R#

Total drag on plate L x B

L  −1/2
  Uo L √
D = 

B τo dx ∼
= 0.664 ρUo2 (BL) ⇒ D ∝ L, D ∝ U 3/2
ν
width 0 

−1/2
ReL

Friction (drag) coefficient:

D ∼ 1.328 1 1
Cf = 1 2 = ⇒ Cf ∝ √ , Cf ∝ √
2
(ρUo ) (BL) R eL L U

Cf
Blasius Laminar Boundary Layer
1.328
Cf ≅
Re L
Turbulent Boundary Layer
0.008

Re L
103 3× 105
Turbulent Boundary Layer
C f for flat plate (JNN 2.3)
⎧⎪ Re x ~ 3×105
Transition at ⎨
⎩⎪Reδ ~ 600

Skin friction coefficient as a function of Re .

A look ahead: Turbulent Boundary Layers

Observe form the previous figure that the function Cf, laminar (Re ) for a laminar
boundary layer is different from the function Cf, turbulent (Re ) for a turbulent boundary
layer for flow over a flat plate.

Turbulent boundary layers will be discussed in proceeding Lecture.

10

4.8 Laminar Boundary Layers for Flow Over a Body of General


Geometry
The velocity profile given in BLBL is the exact velocity profile for a steady, laminar flow
over a flat plate. What is the velocity profile for a flow over any arbitrary body? In general
it is dp/dx = 0 and the boundary layer governing equations cannot be easily solved as was
the case for the BLBL. In this paragraph we will describe a typical approximative procedure
used to solve the problem of flow over a body of general geometry.

1. Solve P-Flow outside B ≡ B0


y
2. Solve boundary layer equations x
(with ∇P term) → get δ ∗ (x)
L ∇P ≠ 0
3. From B0 + δ ∗ → B
U ≠ const.
U

4. Repeat steps (1) to (3) until no

change
B0

• von Karman’s zeroth moment integral equation

τ0 d  2  dU
= U (x)θ(x) + δ ∗ (x)U (x) (4)
ρ dx dx

• Approximate solution method due to Polthausen for general geometry


(dp/dx = 0) using von Karman’s momentum integrals.

The basic idea is the following: we assume an approximate velocity profile (e.g. linear,

4th order polynomial, . . .) in terms of an unknown parameter δ(x). From the velocity

profile we can immediately calculate δ ∗ , θ and τo as functions of δ(x) and the P-Flow

velocity U (x).

Independently from the boundary layer approximation, we obtain the P-Flow solution
outside the boundary layer U (x), dU
dx
.
Upon substitution of δ ∗ , θ, τo , U (x), dU
dx
in von Karman’s moment integral equation(s)
we form an ODE for δ in terms of x.

11

• Example for a 4th order polynomial Polthausen velocity profile


Polthausen profiles - a family of profiles as a function of a single parameter Λ(x)
(shape function factor).

� Assume an approximate velocity profile, say a 4th order polynomial:

u (x, y) y  y 2  y 3  y 4
= a (x) + b (x) + c (x) + d (x) (5)
U (x, 0) δ δ δ δ
There can be no constant term in (5) for the no-slip BC to be satisfied y = 0,
i.e, u(x, 0) = 0.
We use three BC’s at y = δ

u ∂u ∂ 2u
= 1, = 0, = 0, at y = δ (6)
U ∂y ∂y 2
From (6) in (5), we re-write the coefficients a(x), b(x), c(x) and d(x) in terms
of Λ(x)
a = 2 + Λ/6, b = −Λ/2, c = −2 + Λ/2, d = 1 − Λ/6

� To specify the approximate velocity profile Uu(x,y)


(x,0)
in terms of a single unknown
parameter δ we use the x-momentum equation at y = 0, where u = v = 0
  
∂u ∂u ∂U ∂ 2 u  1 dU δ 2 dU δ 2 (x)
u +v =U + ν 2 ⇒b=− ⇒ Λ(x) =
↓ ∂x ↓ ∂y ∂x ∂y y=0 2 dx ν dx ν
0 0 


1 dp
− ρ dx
ν 2bU
2
δ


dU Λ > 0 : favorable pressure gradient
Observe: Λ ∝ ⇒
dx Λ < 0 : adverse pressure gradient
Putting everything together:

u (x, y) y   y 3  y 4

= 2 −2 + +
U (x, 0) δ δ δ
 
dU δ 2 1  y  1  y 2 1  y 3 1  y 4
+ − + −
dx ν 6 δ 2 δ 2 δ 6 δ

12
� Once the approximate velocity profile Uu(x,y)
(x,0)
is given in terms of a single unknown

parameter δ(x), then δ , θ and τo are evaluated

∞   
∗ u 3 1  dU δ 2 
δ = 1− dy = δ −
U 10 120 dx ν
0

δ  
u u 37 1  dU δ 2  1  dU δ 2 2
θ = 1− dy = δ − −
U U 315 945 dx ν 9072 dx ν
0
  
∂u  μU 1  dU δ 2 
τo = μ = 2+
∂y y=0 δ 6 dx ν
Notes:
- Incipient flow (τo = 0) for Λ = −12. However, recall that once the flow is
separated the boundary layer theory is no longer valid.

- For dU
dx
= 0 → Λ = 0 Pohlhausen profile differs from Blasius LBL only by a
few percent.

� After we solve the P-Flow and determine U (x), dU


dx
we substitute everything into
von Karman’s momentum integral equation (4) to obtain

dδ 1 dU d2 U/dx2
= g(δ) + h(δ)
dx U dx dU/dx
where g, h are known rational polynomial functions of δ.
2
This is an ODE for δ = δ(x) where U, dU , d U are specified from the P-Flow
dx dx2
solution.

General procedure:

1. Make a reasonable approximation in the form of (5),


2. Apply sufficient BC’s at y = δ, and the x-momentum at y = 0 to reduce (5)
as a function a single unknown δ,
3. Determine U (x) from P-Flow, and
4. Finally substitute into Von Karman’s equation to form an ODE for δ(x).
Solve either analytically or numerically to determine the boundary layer
growth as a function of x.

13
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 18

13.021 - Marine Hydrodynamics


Lecture 18

4.9 Turbulent Flow – Reynolds Stress


Assume a flow v with a time scale T . Let τ denote a time scale τ << T . We can then
write for each component of the velocity

ui = ūi + ui (1)

where by definition
 τ
1
ūi = ui dt
τ 0
It immediately follows that

∂ ∂ui
u¯i = ui − ūi = ūi − ūi = 0, also ūi =
etc.

∂x
∂x

Substitute Eq. (1) into continuity and average over τ , i.e., take ( )

∂ui ∂ūi ∂ui ∂ūi


=
+ = 0, =⇒ =0
∂xi ∂xi ∂xi ∂xi


0

∂ui ∂ūi ∂u ∂ui


but
= 0 =
+
i , =⇒
=0
∂xi ∂x ∂xi ∂xi
i

0 , just shown

Substitute Eq. (1) into the momentum equations and take ( )


∂ui ∂ui 1 ∂τij 1 ∂p
+ uj = =− + ν∇2 ui
∂t ∂xj ρ ∂xj ρ ∂xi

⎧ 2
∂ui ∂ūi ∂ui ⎨
ν∇2 ui = ν∇ ui
=
+
; similarly
∂t ∂t ∂t


⎩ ∂p ∂ ∂p̄
∂xi
= ∂xi
(p̄ + p ) = ∂xi
etc.

∂ui


∂ ūi ∂ūi ∂u ∂ 
uj ¯j + uj
= u
ui + ui ) = ūj
(¯ + uj + ūj i +uj u
∂xj ∂xj ∂xj ∂xj ∂xj ∂xj i
     
0 0

but from continuity we have


 ∂
  ∂uj
uj u
i = u
u − ui
∂xj ∂xj j i ∂xj

0→by continuity

and thus we finally obtain

∂ūi ∂ūi 1 ∂p ∂  
+ ūj =− + ν∇2 ūi − uu
∂t
∂xj ρ ∂xi ∂xj i j
  
1 ∂
τ
ρ ∂xj ij

∂ūi ∂ūi 1 ∂
Reynolds averaged N-S equation:
+ ūj = τij − ρui uj
∂t ∂xj ρ ∂xj

Reynolds stress:
τRij ≡ −ρui uj

4.10 Turbulent Boundary Layer Over a Smooth Flat Plate


We have already seen that the function of the friction coefficient Cf (ReL ) differs for laminar
and turbulent flows. In this paragraph we will discuss the case of a turbulent boundary
layer.
Following a procedure similar to that for flow past a body of general geometry, we will
use an approximate velocity profile, obtain the P-Flow solution and eventually substitute
everything into von Karman’s momentum integral equation. The velocity profiles used in
practice are either empirical ((1/7)th power) or semi-empirical (logarithmic) laws.

u
y log
Uo

δ 1/7

U y
Uo u log
o δ

4.10.1 (1/7)th Power Velocity Profile Law


Let the velocity profile be determined by the following empirical law
ū  y 1/7
= (2)
Uo δ

where δ = δ(x) is to be determined.


From equation (2) we can obtain directly δ ∗ and θ
δ
δ∗ =
8
7 ∼
θ = δ = 0.0972 δ
72
However, we need to use an additional empirical law to determine the skin friction.
From Blasius’ law of friction for pipes we obtain an expression for τo
 −1/4
τo Uo δ
= 0.0227
ρUo2 ν

3
From P-Flow for flow past a flat plate we have U (x) = U0 = const, and dp/dx = 0
Substituting δ ∗ , θ, τo , Uo into von Karman’s moment equation
 −1/4
τo d Uo δ 7 dδ
= (θ) =⇒ 0.0227 =
ρUo 2 dx ν 72 dx

This is a 1s t order ODE for δ. One BC is required. We assume that the the flow is
tripped at x = 0, i.e., at x = 0 the flow is already turbulent. Further on, we assume
that the turbulent boundary layer starts at x = 0, i.e., δ(0) = 0. It follows that
 −1/5
Uo x δ ∼
δ (x) ∼
= 0.373x =⇒ = 0.373Re−1/5
ν x x

Compare:

Laminar Boundary Layer Turbulent Boundary Layer (1/7th power law)



δ (x) ∝ x δ (x) ∝ x4/5
 νx  4 1/5
δ∗ ∼
= 1.72 Uo
δ ∗ ∼
= 0.047 νx
Uo

Once the profile has been determined we can evaluate the friction drag


D = 0.036 ρUo2 BL Re−1/5
L

Thus, the friction coefficient for turbulent (tripped and/or ReL > 5 × 105 ) flow over
a flat plate is
D
Cf = = 0.073Re−1/5
1 ρU 2 BL L

2 o

4.10.2 Logarithmic Velocity Profile Law


If the velocity profile is determined by the semi-empirical logarithmic velocity pro­
file law, following an approach similar to that for the 1/7th power law, we obtain
Schoenherr’s formula for the friction coefficient

0.242
 = log10 (ReL Cf )
Cf

4.10.3 Summary of Boundary Layer Over a Flat Plate

Laminar BL (Blasius) Turbulent BL (1/7th power law)

δ δ
∝ Re−1/2 ∝ Re−1/5
x x
x x


δ ∗ = 1.72xRe−1/2
x
∝ x δ ∗ = 0.047xRe−1/5
x
∝ x4/5

τo = 0.0227ρUo2 Re−1/4
δ

τo = 0.332ρUo2 Re−1/2
x
τo = 0.02297ρUo2 Re−1/5
x

D = 0.664ρU02 (BL)Re−1/2
L
D = 0.03625ρU02 (BL)Re−1/5
L

D D
Cf ≡ = 1.328Re−1/2 Cf ≡ = 0.0725Re−1/5
ρUo2 (BL) L
ρUo2 (BL) L

For τo , the cross-over is at Rex ∼ 3.4 x 103 , i.e.,


Cf
(τo )laminar > (τo )turbulent for Rex < 3.4 × 103
(τo )laminar ∼ (τo )turbulent for Rex ∼ 3.4 × 103
(τo )laminar < (τo )turbulent for Rex > 3.4 × 103
C f L ~ RL−
1
2

C fT ~ RL−
1
5

~ 0.01
Therefore, for most prototype scales:
ln (RL)
(Cf )turbulent > (Cf )laminar
(τo )turbulent > (τo )laminar RL ~ 1.6 x 104

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 19

13.021 - Marine Hydrodynamics


Lecture 19

Turbulent Boundary Layers: Roughness Effects


So far, we have assumed a ‘hydraulically smooth’ surface. In practice, it is rarely so, due
to fouling, rust, rivets, etc. . . .

Viscous sublayer

Uo

δv
k = characteristic
roughness height

To account for roughness we first define an ‘equivalent sand roughness’ coefficient k (units:
[L]), a measure of the characteristic roughness height.

The parameter that determines the significance of the roughness k is the ratio
k
δ
k
We thus distinguish the following two cases, depending of the value of the ratio δ
on the
actual surface - e.g., ship hull.
1. Hydraulically smooth surface For k < δv << δ, where δv is the viscous sub-layer
thickness, k does not affect the turbulent boundary layer significantly.

k
<< 1 ⇒ Cf  Cf , smooth ⇒ Cf = Cf (ReL )
δ

1
2. Hydraulically rough surface For k >> δ >> δv , the flow will resemble what is
sketched in the following figure.

separation

k
δv

In terms of sand grains: each sand grain can be thought of as a bluff body. The flow,
thus separates downstream of each sand grain. Recalling that drag due to ‘separation’
= form drag >> viscous drag we can approximate the friction drag as the resultant
drag due to the separation behind each sand grain.
k k
>> 1 ⇒ Cf ≡ Cf , rough ⇒ Cf = Cf ( , ReL )
δ L ����
weak dependence

k/l ↑
Cf C D ≠ F (Re L )
k/l = constant

C f rough

RL
C f smooth

Cf , rough has only a weak dependence on ReL , since for bluff bodies CD =
 F (ReL )

In summary The important parameter is k/δ:


k
<< 1 : hydraulically smooth
δ (x)

k
>> 1 : rough
δ (x)
Therefore, for the same k, the smaller the δ, the more important the roughness k.

4.11.1 Corollaries

1. Exactly scaled models (e.g. hydraulic models of rivers, harbors, etc. . . )

k
Same relative roughness: ∼ const for model and prototype
L
� �
k kL k
= ∼ ReL 1/5
δ Lδ L

k
↑ for ReL ↑
δ

For Re model << Re prototype :

� � � �
k k
<
δ m δ p

(Cf )m < (Cf )p

2. Roughness Allowance. Often, the model is hydraulically smooth while the proto­
type is rough. In practice, the roughness of the prototype surface is accounted for
‘indirectly’.

Cf

Uk
= Rk = constant
ν

ΔCf remains
constant with Rl
RL
C f smooth

Uk
• For the same ship (Re same), different k gives different Rek = .
ν
• For a given Rek , the friction coefficient Cf is increased by almost a constant for
Uk
= Rek = const over a wide range of ReL .
ν
• If the model is hydraulically smooth, can we account for the roughness of the
prototype?
Notice that ΔCf = ΔCf (Rek ) has only a weak dependence on ReL . We can
therefore, run an experiment using hydraulically smooth model, and add ΔCf
to the final friction coefficient for the prototype

Cf (ReL ) = Cf smooth + ΔCf (R )


� ��ek�
not (ReL )
Gross estimate: For ships, we typically use ΔCf = 0.0004.

k Rek ∼ R ek
Reality: =� � = =⇒
δ δ/L ReL ReL 4/5
����
∼ReL −1/5
k
↓ as ReL ↑, i.e., ΔCf smaller for larger ReL .
δ

• Hughes’ Method Adjust for ReL dependence of Cfrough .

Cfrough = Cfsmooth (1 + γ) =⇒ ΔCf = γCfsmooth (ReL )

i.e., As ReL ↑, ΔCf ↓.

4
Chapter 5 - Model Testing.
5.1 Steady Flow Past General Bodies
- In general, CD = CD (Re ).

- For bluff bodies

Form drag >> Friction drag ⇒ CD ≈ const ≡ CP (within a regime)

Recall that the form drag (CP ) has only regime dependence on Reynold’s number,
i.e, its NOT a function of Reynold’s number within a regime.

- For streamlined bodies


CD (Re ) = Cf (Re ) + CP

5.1.1 Steps followed in model testing:

(a) Perform an experiment with a smooth model at ReM (ReM << ReS ) and obtain
the model drag CDM .
(b) Calculate CP M = CDM − Cf M (ReM ) = CP S = CP ; CDM measured, Cf M (ReM )
calculated.
(c) Calculate CDS = CP + Cf S (ReS )
(d) Add ΔCf for roughness if needed.

CP measured

CD predicted
Cf (Rship)
Cf (Rm)

calculated

Cf (Rship) calculated
R
Rm Rship

Caution: In an experiment, the boundary layer must be in the same regime (i.e.,
turbulent) as the prototype. Therefore turbulence stimulator(s) must be added.


TBL �
TBL
LBL
CP turbulent regime

U MODEL

Laminar Cf Turbulent Cf

R Turbulent boundary layer


to be triggered here

5.1.2 Drag on a ship hull For bodies near the free surface, the Froude number Fr is
due to wave effects. Therefore CD = CD (Re , Fr ). In general the ra-
important, �
Re gL3
tio = . It is impossible to easily scale both Re and Fr . For example
Fr ν

Re Lm 1 νm gm

= constant and = ⇒ = 0.032 or = 1000!

Fr Lp 10 νp gp

This makes ship model testing seem unfeasible. Froude’s Hypothesis proves to be
invaluable for model testing
calculate measure indirectly
� �� � � �� �
CD (Re , Fr ) ≈ Cf (Re ) + CR (Fr )
� �� � � �� �
Cf for flat plate residual drag
of equivalent wetted area

In words, Froude’s Hypothesis assumes that the drag coefficient consists of two parts,
Cf that is a known function of Re , and CR , a residual drag that depends on Fr num­
ber only and not on Re . Since Cf (Re ) ∼ Cf (Re )flat plate , we need to run experiments
to (indirectly) get CR (Fr ).

Thus, for ship model testing we require Froude similitude to measure CR (Fr ), while
Cf (Re ) is estimated theoretically.

6
5.1.3 OUTLINE OF PROCEDURE FOR FROUDE MODEL TESTING

(S ≡ ‘SHIP’ M ≡‘MODEL’; in general νS  = νM , and ρS =



ρM )

1. Given US , calculate: FrS = US / gLS = FrM


2. For Froude similitude, tow model at: UM = FrS gLM

3. Measure total resistance (drag) of model: Measure DM

DM
4. Calculate total drag coefficient for model: CDM = 2
0.5ρM UM SM
����
wetted area

0.075
5. Use ITTC line to calculate Cf (ReM ): Cf (ReM ) =
(log10 ReM − 2)2

6. Calculate residual drag of model: CRM = CDM − Cf (ReM )

7. Froude’s Hypothesis: CRM (ReM , Fr ) = CRM (Fr ) = CRS (Fr ) = CR (Fr )

0.075
8. Use ITTC line to calculate Cf (ReS ): Cf (ReS ) =
(log10 ReS − 2)2

9. Calculate total drag coefficient for ship: CDS = CR (Fr ) + Cf (ReS ) + ΔC


����f

= 0.0004
typical value

� �
10. Calculate the total drag of ship: DS = CDS · 0.5ρS US2 SS
����
wetted area

11. Calculate the power for the ship: PS = DS US

12. Repeat for a series of US

7
13.021 – Marine Hydrodynamics, Fall 2004

Lecture 20

13.021 - Marine Hydrodynamics


Lecture 20

Chapter 6 - Water Waves


6.1 Exact (Nonlinear) Governing Equations for
Surface Gravity Waves, Assuming Potential Flow

Free surface definition


y = η ( x, z , t ) or F ( x, y, z , t ) = 0
y

x
z

B(x, y, z,t) = 0

Unknown variables


Velocity field:
v (x, y, z, t) = ∇φ (x, y, z, t)
Position of free surface:
y = η (x, z, t) or F (x, y, z, t) = 0
Pressure field:
p (x, y, z, t)

Governing equations
Continuity:
∇2 φ = 0 y < η or F < 0
Bernoulli for P-Flow:
∂φ
∂t
+ 12 |∇φ|2 + p−pa
ρ
+ gy = 0; y < η or F < 0
Far way, no disturbance:
∂φ/∂t, ∇φ → 0 and p = pa − ρgy
 
atmospheric hydrostatic

1
Boundary Conditions

1. On an impervious boundary B (x, y, z, t) = 0, we have KBC:


∂φ  (x, t) · n
v · n
ˆ = ∇φ · n
ˆ= =U ˆ (x, t) = Un on B = 0
∂n

Alternatively: a particle P on B remains on B, i.e., B is a material surface. For


example if P is on B at t = t0 , P stays on B for all t.

B(xP , t0 ) = 0, then B(xP (t), t) = 0 for all t,

so that, following P B is always 0.


DB ∂B
∴ = + (∇φ · ∇) B = 0 on B = 0
Dt ∂t

For example, for a flat bottom at y = −h ⇒ B = y + h = 0 ⇒


  
DB ∂φ ∂ ∂φ
= (y + h) = 0 ⇒ = 0 on B = y + h = 0
Dt ∂y ∂y ∂y
  
=1

2. On the free surface, y = η or F = y − η(x, z, t) = 0 we have KBC and DBC.

KBC: free surface is a material surface, no normal velocity relative to the free surface.
A particle on the free surface remains on the free surface for all times.
DF D ∂φ ∂η ∂φ ∂η ∂φ ∂η
=0= (y − η) = − − − on y = η
Dt Dt ∂y ∂t ∂x 
∂x ∂z 
∂z 
 still
vertical slope slope
unknown
velocity of f.s. of f.s.

DBC: p = pa on y = η or F = 0. Apply Bernoulli equation at y = η:

∂φ 1
+ |∇φ|2 + g η = pa on y = η
∂t 2   
still unknown
non-linear term

2
6.2 Linearized (Airy) Wave Theory
Assume small wave amplitude compared to wavelength, i.e., small free surface slope
A
<< 1
λ

Wave height H crest


Wave amplitude A
H = A/2

SWL

Water depth h
trough wavelength Wave period T
λ

Consequently
φ η
, << 1
λ2 /T λ
We keep only linear terms in φ, η.


For example: ()|y=η = ()y=0 + η ()| + . . . Taylor series
    ∂y  y=0
keep
discard

6.2.1 BVP In this paragraph we state the Boundary Value Problem for linear (Airy) waves.

∂ 2φ ∂φ
+g =0
∂t 2 ∂y
y=0
∇ 2φ = 0
y = -h

∂φ
=0
∂y

Finite depth h = const Infinite depth


GE: ∇2 φ = 0, −h < y < 0 ∇2 φ = 0, y < 0
∂φ
BKBC: ∂y = 0, y = −h ∇φ → 0, y → −∞

FSKBC: ∂φ ∂η ⎬
∂y = ∂t , y = 0 2
→ ∂∂t2φ + g ∂φ
∂y = 0
FSDBK: ∂φ
∂t + gη = 0, y = 0 ⎭

Introducing the notation {} for infinite depth we can rewrite the BVP:

Constant finite depth h {Infinite depth}


2
∇2 φ = 0, −h < y < 0 ∇ φ = 0, y < 0 (1)
∂φ
= 0, y = −h {∇φ → 0, y → −∞} (2)
∂y
2 
∂2φ ∂φ ∂ φ ∂φ
+ g = 0, y = 0 +g = 0, y = 0 (3)
∂t2 ∂y ∂t2 ∂y

Given φ calculate:
   
1 ∂φ  
1 ∂φ 
η (x, t) = − η (x, t) = − (4)
g ∂t y=0 g ∂t y=0

∂φ ∂φ
p − pa = −ρ − ρgy p − pa = −ρ − ρgy (5)
 ∂t   ∂t 
hydrostatic hydrostatic
dynamic dynamic

4
6.2.2 Solution Solution of 2D periodic plane progressive waves, applying separation of
variables.
We seek solutions to Equation (1) of the form eiωt with respect to time. Using the
KBC (2), after some algebra we find φ. Upon substitution in Equation (4) we can
also obtain η.

gA cosh k (y + h) gA ky
φ= sin (kx − ωt) φ= sin (kx − ωt) e
ω cosh kh ω
⎧ ⎫
⎨ ⎬
η 
= A cos (kx − ωt) = A cos (kx − ωt)
η 
⎩ ⎭
using (4) using(4)

where A is the wave amplitude A = H/2.


Exercise Verify that the obtained values for φ and η satisfy Equations (1), (2), and
(4).
6.2.3 Review on plane progressive waves

(a) At t = 0 (say), η = A cos kx → periodic in x with wavelength: λ = 2π/k


Units of λ : [L]
λ
k = wavenumber = 2 /λ [L-1]
x K

(b) At x = 0 (say), η = A cos ωt → periodic in t with period: T = 2π/ω


Units of T : [T ]
T

ω = frequency = 2 /T [T-1], e.g. rad/sec


t
    
(c) η = A cos k x − ωk t ω L
Units of :
k T
ω ω 
Following a point with velocity , i.e., xp = t + const, the phase of η does
k k
ω λ
not change, i.e., = ≡ Vp ≡ phase velocity.
k T

5
6.2.4 Dispersion Relation
So far, any ω, k combination is allowed. However, recall that we still have not made
use of the FSBC Equation (3). Upon substitution of φ in Equation (3) we find that
the following relation between h, k, and ω must hold:

∂ 2φ ∂φ
+ g =0 −→ −ω 2 cosh kh + gk sinh kh = 0 ⇒ ω 2 = gk tanh kh
∂t2 ∂y ↑
φ= gA
ω
sin(kx−ωt)f (z)

• This is the Dispersion Relation



ω 2 = gk tanh kh ω 2 = gk (6)

Given h, the Dispersion Relation (6) provides a unique relation between ω and k,
i.e., ω = ω(k; h) or k = k(ω; h).
• Proof

C
ω2h
C ≡ = (kh) tanh (kh)
g 
kh

from (6)
1 tanh kh
C
= tanh kh
kh
kh → obtain unique solution for k
kh =f(c)

• Comments
- General As ω ↑ then k ↑, or equivalently as T ↑ then λ ↑.
  
λ ω g g
- Phase speed Vp ≡ = = tanh kh Vp =
T k k k
Therefore as T ↑ or as λ ↑, then Vp ↑, i.e., longer waves are ‘faster’ in terms of
phase speed.
- Water depth effect For waves the same k (or λ), at different water depths, as h ↑
then Vp ↑, i.e., for fixed k Vp is fastest in deep water.
- Frequency dispersion Observe that Vp = Vp (k) or Vp (ω). This means that waves of
different frequencies, have different phase speeds, i.e., frequency dispersion.

6
6.2.5 Solutions to the Dispersion Relation : ω 2 = gk tanh kh
Property of tanh kh:
long waves
shallow water
  
sinh kh 1 − e−2kh ∼ kh for kh << 1. In practice h < λ/20
tanh kh = = =
cosh kh 1 + e−2kh 1 for kh >∼ 3. In practice h > λ2
  
short waves
deep water

Shallow water waves Intermediate depth Deep water waves


or long waves or wavelength or short waves
kh << 1 Need to solve ω 2 = gk tanh kh kh >> 1
∼ h < λ/20 given ω, h for k ∼ h > λ/2
(given k, h for ω - easy!)

ω2 ∼
= gk · kh → ω = gh k (a) Use tables or graphs (e.g.JNN fig.6.3) ω 2 = gk
√ g 2
λ = gh T ω 2 = gk tanh kh = gk∞ λ= T

k∞ λ Vp  
⇒ = = = tanh kh λ(in ft.) ≈ 5.12T 2 (in sec.)
k λ∞ Vp∞
(b) Use numerical approximation
(hand calculator, about 4 decimals )

i. Calculate C = ω 2 h/g
ii. If C > 2: ”deeper” ⇒
kh ≈ C(1 + 2e−2C − 12e−4C + . . .)
If C < 2: ”shallower” ⇒

kh ≈ C(1 + 0.169C + 0.031C 2 + . . .)
No frequency dispersion Frequency dispersion Frequency dispersion
 
 g g
Vp = gh Vp = tanh kh Vp = λ
k 2π

6.3 Characteristics of a Linear Plane Progressive Wave

λ η(x,t) = y
2π Vp
k= A
λ MWL x

ω=
T

H = 2A
h

Define U ≡ ωA

Linear Solution:
Ag cosh k (y + h)
η = A cos (kx − ωt) ; φ= sin (kx − ωt) , where ω 2 = gk tanh kh
ω cosh kh
6.3.1 Velocity field

Velocity on free surface v (x, y = 0, t)


1 ∂η
u(x, 0, t) ≡ Uo = Aω cos (kx − ωt) v(x, 0, t) ≡ Vo = Aω sin (kx − ωt) =
tanh kh ∂t

Velocity field v (x, y, t)

∂φ Agk cosh k (y + h) ∂φ Agk sinh k (y + h)


u= = cos (kx − ωt) v= = sin (kx − ωt)
∂x ω cosh kh ∂y ω cosh kh
cosh k (y + h) sinh k (y + h)

=  cos (kx − ωt) ⇒ Aω
=  sin (kx − ωt) ⇒
sinh kh sinh kh
U U

⎧ ⎧

⎨ ⎪
∼ eky sinh k (y + h) ⎨ ∼ e
ky
u cosh k (y + h) deep water v deep water
= =
Uo cosh kh ⎪
⎩ ∼1 Vo sinh kh ⎪
⎩ ∼1+ y
shallow water h
shallow water

• u is in phase with η • v is out of phase with η

Velocity field v (x, y)


Shallow water Intermediate water Deep water

6.3.2 Pressure field

• Total pressure p = pd − ρgy.


• Dynamic pressure pd = −ρ ∂φ
∂t
.
• Dynamic pressure on free surface pd (x, y = 0, t) ≡ pdo

Pressure field

Shallow water Intermediate water Deep water

cosh k (y + h)
pd = ρgη pd = ρgA cos (kx − ωt) pd = ρgeky η
cosh kh
cosh k (y + h)
= ρg η
cosh kh
pd
same picture as Uuo
p do
pd (−h) pd (−h) 1 pd (−h)
= 1 (no decay) = = e−ky
p do pdo cosh kh p do
 
p= ρg(η − y) p = ρg ηeky − y
  
“hydrostatic” approximation

V p = gh V p = gλ

y y y y y
y
x x
∇ ∇

kh << 1 kh >> 1

pd p ( − h) pd o pd o p (− h) pd o
p (− h) p (− h)

Pressure field in shallow water Pressure field in deep water

10

6.3.3 Particle Orbits (‘Lagrangian’ concept)


Let xp (t), yp (t) denote the position of particle P at time t.
Let (x̄; ȳ) denote the mean position of particle P.
The position P can be rewritten as xp (t) = x̄ + x (t), yp (t) = ȳ + y  (t), where
(x (t), y  (t)) denotes the departure of P from the mean position.

In the same manner let v ≡ v(¯ x, y,


¯
t) denote the velocity at the mean position and
vp ≡ v (xp , yp , t) denote the velocity at P.

P (x , y ) vp = v (x̄ + x , ȳ + y  , t) =⇒
P P
(x' , y ' ) TSE

∂v ∂v
vp = v (¯
x, y,
¯ t) + ¯ t) x  +
x, y,
(¯ ¯ t) y  + . . . ⇒
(x̄, y,
(x, y) ∂x ∂y
  
ignore - linear theory

vp ∼
= v

To estimate the position of P, we need to evaluate (x (t), y  (t)):



 cosh k (ȳ + h)
x = dt u (¯
x, y,
¯ t) = dt ωA cos (kx¯ − ωt) ⇒
sinh kh
cosh k (ȳ + h)
= −A sin (kx̄ − ωt)
sinh kh

 sinh k (ȳ + h)
y = dt v (¯
x, y,
¯ t) = dt ωA sin (kx̄ − ωt) ⇒
sinh kh
sinh k (ȳ + h)
= A cos (kx̄ − ωt)
sinh kh

Check: On ȳ = 0, y  = A cos (kx̄ − ωt) = η, i.e., the vertical motion of a free surface
particle (in linear theory) coincides with the vertical free surface motion.
It can be shown that the particle motion satisfies

x2 y 2 (xp − x̄)2 (yp − ȳ)2


+ = 1 ⇔ + =1
a2 b2 a2 b2
cosh k (ȳ + h) sinh k (ȳ + h)
where a = A and b = A , i.e., the particle orbits form
sinh kh sinh kh
closed ellipses with horizontal and vertical axes a and b.

11
crest
Vp

ky
A
(a) deep water kh >> 1: a = b = Ae A
ky
circular orbits with radii Ae decreasing
exponentially with depth
trough
ky
Ae

A
Vp = gh
(b) shallow water kh << 1:

A y
a= = const. ; b = A(1+ )
kh h

decreases linearly

with depth A/kh

Vp

(c) Intermediate depth

P Vp
A
Q S Q S

R R R

λ
6.3.4 Summary of Plane Progressive Wave Characteristics

f (y) Deep water/ short waves Shallow water/ long waves

kh > π (say) kh << 1

cosh k(y+h)
cosh kh
= f1 (y) ∼ eky 1
e.g.pd

1
cosh k(y+h)
sinh kh
= f2 (y) ∼ eky kh
e.g.u, a

y
sinh k(y+h)
sinh kh
= f3 (y) ∼ eky 1+ h
e.g. v, b

13

C (x) = cos (kx − ωt) S (x) = sin (kx − ωt)

(in phase with η) (out of phase with η)

η
A
= C (x)

u v

= C (x) f2 (y) Aω
= S (x) f3 (y)

pd
ρgA
= C (x) f1 (y)

y� x�
A
= C (x) f3 (y) A
= −S (x) f2 (y)

a
A
= f2 (y) b
A
= f3 (y)

b
a

14

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 21

13.021 - Marine Hydrodynamics


Lecture 21

6.4 Superposition of Linear Plane Progressive Waves


1. Oblique Plane Waves

v
k
kz
z
kx
v
Vp k = (k x , k z )

θ
x

(Looking up the y-axis from


below the surface)

Consider wave propagation at an angle θ to the x-axis

k·x

  
η =A cos(kx cos θ + kz sin θ −ωt) = A cos (kx x + kz z − ωt)
gA cosh k (y + h)
φ= sin (kx cos θ + kz sin θ − ωt)
ω cosh kh 
ω 2 =gk tanh kh; kx = k cos θ, kz = k sin θ, k = kx2 + kz2

2. Standing Waves

+ Same A, k, ω, no phase shift


y

η =A cos (kx − ωt) + A cos (−kx − ωt) = 2A cos kx cos ωt


2gA cosh k (y + h)
φ=− cos kx sin ωt
ω cosh kh

90o at all times


y
t = 0, T, 2T, …

amplitude
2A

node
T 3T
t= , ,L T 3T 5T
2 2
antinode
t= , , L

4 4 4

∂η ∂φ nπ nλ
∼ = · · · sin kx = 0 at x = 0, =
∂x ∂x k 2

∂φ

Therefore,
= 0. To obtain a standing wave, it is necessary to have perfect

∂x

x=0
reflection at the wall at x = 0.
AR
Define the reflection coefficient as R ≡ (≤ 1).
AI

A I = AR
x AR
R= =1
AI

3. Oblique Standing Waves

ηI =A cos (kx cos θ + kz sin θ − ωt)


ηR =A cos (kx cos (π − θ) + kz sin (π − θ) − ωt)

z
ηR
θR
θ θθI
x θ R = π − θI
θ
ηI

Note: same A, R = 1.

x k x z k z−ωt
     
ηT = ηI + ηR = 2A cos (kx cos θ) cos (kz sin θ − ωt)
    
standing wave in x propagating wave in z

and

2π 2π ω
λx = ; VPx = 0; λz = ; VPz =
k cos θ k sin θ k sin θ
Check:

∂φ ∂η
∼ ∼ · · · sin (kx cos θ) = 0 on x = 0
∂x ∂x

4. Partial Reflection

ηI ηR


ηI =AI cos (kx − ωt) = AI Re ei(kx−ωt)

ηR =AR cos (kx + ωt + δ) = AI Re R e−i(kx+ωt)

R: Complex reflection coefficient


AR
R = |R| e−iδ , |R| =
AI


ηT =ηI + ηR = AI Re ei(kx−ωt) 1 + Re−2ikx

|ηT |2 =A2I 1 + |R|2 + 2 |R| cos (2kx + δ)

free surface
| ηT | wave envelope
AI λ
2
1+ | R |2
∇ x

node
antinode
At node,
|ηT | = |ηT |min = AI (1 − |R|) at cos (2kx + δ) = −1 or 2kx + δ = (2n + 1) π
At antinode,
|ηT | = |ηT |max = AI (1 + |R|) at cos (2kx + δ) = 1 or 2kx + δ = 2nπ

λ
2kL = 2π so L =
2

|ηT |max − |ηT |min


|R | = = |R (k)|
|ηT |max + |ηT |min

4
5. Wave Group
2 waves, same amplitude A and direction, but ω and k very close to each other.

VP1


η1 = Aei(k1 x−ω1 t)


η2 = Aei(k2 x−ω2 t)
VP2
ω1,2 =ω1,2 (k1,2 ) and VP1 ≈ VP2


ηT = η1 + η2 =  Aei(k1 x−ω1 t) 1 + ei(δkx−δωt) with δk = k2 − k1 and δω = ω2 − ω1

Vg

λg =
δk
2A
VP1 ≈ VP2

2π 2π
T= Tg =
ω δω

= λ1 ≈ λ 2

k1


|ηT |max = 2 |A| when δkx − δωt = 2nπ ⎬ δω
xg = Vg t, δkVg t−(δω) t = 0 then Vg =
⎭ δk
|ηT |min = 0 when δkx − δωt = (2n + 1) π

In the limit,

dω 
δk, δω → 0, Vg = ,

dk

k1 ≈k2 ≈k

and since

ω 2 = gk tanh kh ⇒

ω  1  2kh


Vg = 1+
k2
  sinh 2kh
 
 

Vp n


(a) deep water kh >> 1 ⎪




⎪ Vg
n= Vg
= 1 ⎪

Vp 2 ⎪



(b) shallow water kh << 1 ⎬
VP
Vg ≤ Vp
n= Vg
= 1 (no dispersion) ⎪


Vp ⎪



(c) intermediate depth ⎪



Appear Disappear
1 ⎪

2
<n<1

6.5 Wave Energy - Energy Associated with Wave Motion.


For a single plane progressive wave:

Energy per unit surface area of wave


• Potential energy PE • Kinetic energy KE
0 η
PE without wave = ρgydy = − 12 ρgh2 KEwave = dy 12 ρ (u2 + v 2 )
−h −h

PE with wave ρgydy = 12 ρg (η 2 − h2 ) Deep water = · · · = 1
ρgA2 to leading order
4  
−h KE const in x,t
1 1
PEwave = 2
ρgη 2 = 2
ρgA2 2
cos (kx − ωt) Finite depth = · · ·
Average energy over one period or one wavelength
PEwave = 14 ρgA2 KEwave = 14 ρgA2 at any h

• Total wave energy in deep water:



E = PE + KE = 12 ρgA2 cos2 (kx − ωt) + 12
• Average wave energy E (over 1 period or 1 wavelength) for any water depth:
E = 12 ρgA2 [ 1
2
+ 1
2
] = 12 ρgA2 = Es ,
↑ ↑
PE KE

Es ≡ Specific Energy: total average wave energy per unit surface area.
• Linear waves: PE = KE = 12 Es Vp
(equipartition).
x
• Nonlinear waves: KE > PE.
Vp
E
Es PE = Es cos2 (kx − ωt)
1 PE
PE = 12 E

½ KE
KE = 1
2
E =
x

1
Recall: cos2 x = 2
+ 12 cos 2x

6.6 Energy Propagation - Group Velocity

S
Vp

E = E s per area V

Consider a fixed control volume V to the right of ‘screen’ S. Conservation of energy:

dW dE
= = J­
dt
 dt
 
rate of work done on S rate of change of energy in V energy flux left to right

where

η  
dφ ∂φ
J- = pu dy with p = −ρ + gy and u =
dt ∂x
−h

1 ω  1

2 2kh

J- = 2 ρgA 1 + = E (nVp ) = EVg
k  2
     sinh 2kh
 
E n
Vp
  
Vg

e.g. A = 3m, T = 10 sec → J- = 400KW /m

6.7 Equation of Energy Conservation

Δx
1 2
x

F1 1 F2 2 E = E (x ), F = F ( x )

h = h(x)

J-1 − J-2 Δt = ΔEΔx



∂J­

J-2 = J-1 + Δx +
· · ·

∂x

1
∂E ∂J­
+ = 0, but J- = Vg E
∂t ∂x
∂E ∂

+ Vg E = 0
∂t ∂x

∂E

1. = 0, Vg E = constant in x for any h(x).


∂t
2. Vg = constant (i.e., constant depth, δk << k)
 
∂ ∂
+ Vg E = 0, so E = E (x − Vg t) or A = A (x − Vg t)
∂t ∂x

i.e., wave packet moves at Vg .

6.8 Steady Ship Waves, Wave Resistance

D
U
Vp = U
2A

E = 21 ρgA2
E = 0 ahead of ship

(
F = Vg E = ( 12 U ) 12 ρgA2 )
L

x=0

C.V.

• Ship wave resistance drag Dw


Rate of work done = rate of energy increase
d

Dw U + J- = EL = EU
dt
deep water

1   
Dw = (EU − EU 2 ) = 12 E = 14 ρgA2 ⇒ Dw ∝ A2
force / length U energy / area

• Amplitude of generated waves


The amplitude A depends on U and the ship geometry. Let ≡ effective length.
L
-+ +-
lL
To approximate the wave amplitude A superimpose a bow wave (ηb ) and a stern wave
(ηs ).
ηb = a cos (kx) and ηs = −a cos (k (x + ))
ηT = ηb + ηs


A = |ηT |max = 2a sin 21 k  ← envelope amplitude



Dw = 14 ρgA2 = ρga2 sin2 12 k ⇒ Dw = ρga2 sin2 12 Ug2

• Wavelength of generated waves To obtain the wave length, observe that the phase
speed of the waves must equal U . For deep water, we therefore have

ω deep g U2
Vp = U ⇒ = U water −→ = U , or λ = 2π
k k g

10
• Summary Steady ship waves in deep water.

U = ship speed

g g U2
Vp = = U ; so k = 2 and λ = 2π
k U g
L =ship length, ∼ L
   


2 1 g ∼ 1 1
2 2
Dw =ρga sin 2 U 2 = ρga sin 2 ∼ 2
= ρga sin 2
2Fr2L 2Fr2L

1
Fl = ≈ 0.56 ⇒
π
max at:
Dw
ρga 2 U hull ≈ 0.56 gl ≅ 0.56 gL ⇒ U hull ∝ L

U
Fl = , where l ≤ L
0 gl
1 Increasing U
π

Small speed U
• Short waves
• Significant wave cancellation
• Dw ~ small

11

13.021 – Marine Hydrodynamics, Fall 2004

Lecture 22

13.021 - Marine Hydrodynamics


Lecture 22

6.9 Wave Forces on a Body

UP

U = ωA
U ωA
Re = =
ν ν
UT AωT A
Kc = = = 2π
  

 
F A  h
CF = =f , , Re , , roughness, . . .
ρgA2 λ
 λ
 λ
Wave Diffraction
steepness parameter

6.9.1 Types of Forces

1. Viscous forces Form drag, viscous drag = f (Re , Kc , roughness, . . .).

• Form drag (CD )


Associated primarily with flow separation - normal stresses.

�����
��������������

• Friction drag (CF )



Associated with skin friction τw , i.e., F ∼ τw dS.
body
(wetted surface)

τω b.l.

2. Inertial forces Froude-Krylov forces, diffraction forces, radiation forces.


Forces arising from potential flow wave theory,


∂φ 1
F = pˆ
ndS, where p = −ρ + gy + |∇φ|2
∂t 2  
body
(wetted surface) =0, for linear theory,
small amplitude waves

For linear theory, the velocity potential φ and the pressure p can be decomposed to

φ= φI + φD + φR
  
Incident wave Diffracted wave Radiated wave
potential (a) potential (b.1) potential (b.2)
p ∂φI ∂φD ∂φR
− = + + + gy
ρ ∂t ∂t ∂t

(a) Incident wave potential

• Froude-Krylov Force approximation When  << λ, the incident wave field is not
significantly modified by the presence of the body, therefore ignore φD and φR .
Froude-Krylov approximation:
 
φ ≈ φI ∂φ
⇒ F F K =
I

∂φI −ρ + gy ndS
ˆ ← wave
can calculate knowing (incident)
kinematics (and body geometry)
p ≈ −ρ + gy ∂t
∂t body   
surface
≡ pI

• Mathematical approximation After


 the divergence theorem, the F F K
applying 
can be rewritten as F F K = − pI ndS
ˆ =− ∇pI d∀.
body body
surface volume

If the body dimensions are very small comparable to the wave length, we can
assume that ∇pI is approximately constant through the body volume ∀ and
‘pull’ the ∇pI out of the integral. Thus, the F F K can be approximated as
   

F F K ∼
= − ∇pI d∀ = 
∀ − ∇pI
at body body
at body
center body center
volume volume

The last relation is particularly useful for small bodies of non-trivial geometry ­
for 13.021, that is all bodies that do not have a rectangular cross section.

3
(b) Diffraction and Radiation Forces

(b.1) Diffraction or scattering force When  ≮ < λ, the wave field near the body
will be affected even if the body is stationary, so that no-flux B.C. is satisfied.


� φ��
φ�� ���������������

� φ�� φ��
� φ��
∂φ ∂
� φ�� =�= (φ � + φ � )
∂� ∂�

∂φ � ∂φ
� �� = − � ← �����
∂� ∂�
�  
∂φD
F D = −ρ ndS
ˆ
∂t
body
surface

(b.2) Radiation Force - added mass and damping coefficient Even in the
absence of an incident wave, a body in motion creates waves and hence inertial wave
forces.

��


φ��
� φ�� �
� � ∂φ � �
= �⋅�
� ∂�


 
∂φR
F R = −ρ ndS
ˆ = − mij U̇j − dij Uj
∂t  
body added wave
surface mass radiation
damping

6.9.2 Important parameters


(1)Kc = UT
= 2π A ⎪

 ⎪

Interrelated through maximum wave steepness

A
≤ 0.07

 ⎪

A 
(2)diffraction parameter
λ  λ
≤ 0.07

• If Kc ≤ 1: no appreciable flow separation, viscous effect confined to boundary layer


(hence small), solve problem via potential theory. In addition, depending on the value
of the ratio λ ,

– If λ << 1, ignore diffraction , wave effects in radiation problem (i.e., dij ≈


0, mij ≈ mij infinite fluid added mass). F-K approximation might be used,
calculate F F K .

– If λ
>> 1/5, must consider wave diffraction, radiation ( A ≤ 0.07
/λ
≤ 0.035).

• If Kc >> 1: separation important, viscous forces can not be neglected. Further on if


 0.07 
≤ so << 1 ignore diffraction, i.e., the Froude-Krylov approximation is valid.
λ A/ λ
1
F = ρ2 U (t) |U (t)|CD (Re )
2 
relative
velocity

• Intermediate Kc - both viscous and inertial effects important, use Morrison’s formula.
1
F = ρ2 U (t)|U (t)|CD (Re ) + ρ3 U˙ Cm (Re , Kc )
2

• Summary

I
Limiting case:
wave breaking occurs

II III

I. Use: CD and F − K approximation.

II. Use: CF and F − K approximation.

III. CD is not important and F − K approximation is not valid.

Anda mungkin juga menyukai