You are on page 1of 27

Accepted Manuscript

Generalized Kutta-Joukowski theorem for multi-vortex and multi-airfoil flow


with vortex production (general model)

Bai ChenYuan, Li Juan, Wu ZiNiu

PII: S1000-9361(14)00046-6
DOI: http://dx.doi.org/10.1016/j.cja.2014.03.014
Reference: CJA 256

To appear in:

Received Date: 22 July 2013


Revised Date: 13 September 2013
Accepted Date: 13 October 2013

Please cite this article as: B. ChenYuan, L. Juan, W. ZiNiu, Generalized Kutta-Joukowski theorem for multi-vortex
and multi-airfoil flow with vortex production (general model), (2014), doi: http://dx.doi.org/10.1016/j.cja.
2014.03.014

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Generalized Kutta-Joukowski theorem for multi-vortex
and multi-airfoil flow with vortex production (general
model)
*
BAI ChenYuan, LI Juan, WU ZiNiu
a
School of Aerospace , Tsinghua Universty, Beijing 100084, China

Received 22 July 2013; in revised form 13 Oct 2013; accepted 13 Oct 2013

Abstract
By using a special momentum approach and with the help of interchange between singularity velocity and
induced flow velocity, we derive in a physical way explicit force formulas for two-dimensional inviscid flow
involving multiple bound and free vortices, multiple airfoils and vortex production. These force formulas hold
individually for each airfoil thus allowing for force decomposition and the contributions to forces from
singularities (such as bound and image vortices, sources and doublets) and bodies out of an airfoil are related to
their induced velocities at the location of singularities inside this airfoil. The force contribution due to vortex
production is related to the vortex production rate and the distance between each pair of vortices in production,
thus frame-independent. The formulas are validated against a number of standard problems. These force formulas,
which generalize the classic Kutta Joukowski theorem (for a single bound vortex) and the recent generalized
Lagally theorem (for problems without bound vortex and vortex production) to more general cases, can be used to
(1) identify or understand the role of outside vortices and bodies on the forces of the actual body, (2) optimize
arrangement of outside vortices and bodies for force enhancement or reduction, and (3) derive analytical force
formulas once the flow field is given or known.

Keywords: Drag force; Lift force; multibody; multiple vortices; vortex production.

1. Introduction1

In the classic Kutta Joukowski theorem for steady potential flow around a single airfoil, the lift is
related to the circulation of bound vortex1. The circulation of the bound vortex is determined by the
Kutta condition, due to which the role of viscosity is implicitly incorporated though explicitly ignored2.
The lift predicted by Kutta Joukowski theorem within the framework of inviscid flow theory is quite
accurate even for real viscous flow, provided the flow is steady and unseparated, see Anderson3 for
more details.
The Kutta Joukowski theorem does not hold for problems with free vortices or other bodies outside
of the body. These problems have been attracting great attentions since more than three decades ago,
due to their wide applications in unsteady flows4-6 and in multibody flows such as multi-turbine flow7,
multi-blade flow8, multi-element airfoil flow9, multi-wing aerodynamics as for dragonfly10, and flows
in staggered cylinders11. Early studies considering the interaction of a single vortex with a wing can be
found in Saffman12. Streitlien and Triantafyllou13 derived force formulas for a single Joukowski airfoil
surrounded with point vortices convected freely.
Traditionally, force formulas for complex flows are either in algebraic form related to speed and
strength of singularities (called singularity approach here)14-Error! Reference source not found. or in integral form
(called integral approach here)12,19-20.
The force formulas by the singularity approach are basically worked out through using the complex
potential theory and the unsteady Blasius theorem21. The forces are expressed in terms of the speeds of
both real and image vortices, plus the time variation of an additional integral term representing all the

*Corresponding author. Tel.: +86-10-62784116.


E-mail address: ziniuwu@tsinghua.edu.cn

1
effects other than the motion of vortices outside of the body.
In the integral approach, force formulas are derived directly from the basic equations of the flow,
notably from equations of vortex dynamics. Forces may be expressed as volume integral of vorticity
moments in the whole space including the solid body12,19. With the help of auxiliary flow potentials,
Howe20 derived an integral approach with a surface integral for the viscous force but with the volume
integral only defined for the space occupied by the fluid.
The main disadvantages of these force formulas are that they do not allow for direct decomposition
of forces according to bodies for multibody flow, and the contribution from each flow structure are not
easily identifiable. An exception is the integral approach of Howe20 which has been recently extended
to multiple bodies by Ragazzo & Tabak22 and Chang, Yang & Chu23 through the introduction of
auxiliary functions, the resulting force formulas are still not in explicit form. Multibody problem is
difficult as stated by Crowdy11 who derived a general method for finding the potential flow solution for
staggered cylinders with given bound vortices. For two parallel plates in a uniform flow Sakajo24
successfully obtained potential flow solution and vortex stationary position. Forces are integrated when
the vortex is at the stationary position.
In some other works, the force contribution from outside vortex or doublet is expressed as the
induced velocity either on the location of outside singularity21 or inside the body9. For the case of an
airfoil interacting with one outside vortex, Katz and Plotkin9 expressed the force in terms of the
induced velocity at the body center. The role of real vortex outside of the body is represented by the
induced velocity at the body center. This result was obtained under the lumped vortex assumption and
extended to the case of multi-airfoil and multi-vortex flow by Bai&Wu25 still under lumped vortex
assumption. The presence of singularities such as sources and doublets outside of a body has been also
studied in the framework of Lagally theorem21,25. Wu, Yang &Young27 extended the Lagally theorem to
the case of two dimensional flow with multiple bodies moving in a still fluid in the presence of multiple
free vortices. Bound vortex and vortex production are not considered in this work and the force
formulas are extended to multiple bodies without proof.
In this paper, we consider two dimensional potential flow with multiple bodies and multiple free
vortices. Moreover, we include both bound vortices and vortex production, which are however not
considered in the work of Wu, Yang &Young27. The purpose is to derive force formulas which hold
individually for each airfoil and which are explicit allowing for easy identification of the role of each
outside vortices and bodies. Due to vortex production on the surface of the body, the unsteady Blasius
equation given by Thomson21 and used by Wu, Yang &Young27 cannot be directly applied here since it
is based on irrotational assumption on the body. Hence we will use an integral approach based on
specially devised control volumes to find the force formulas.
This paper will be organized as follows. In section 2 we first use a momentum approach to relate the
lift force and induced drag force to the speeds of singularities inside and outside of a single body
(singularity velocity method). We then relate the force terms due to the outside singularities to the
induced velocities inside the body and express the forces in terms of the relative induced velocities and
strengths of singularities inside of the body (induced velocity method). The induced velocity method is
then extended to the case of multibody flow.
In section 3, we use several standard cases to validate the force formulas. These include the problem
of a standing vortex pair behind a circular cylinder, a source doublet outside of a circular cylinder, the
problem of two circular cylinders with given bound vortices, for which Crowdy11 gives exact solution.
The last example, which is used to test the singularity approach and the force formula for vortex
2
production, is the drag problem for Karman vortex street, a very difficulty problem since the drag is not
related to the shape of the body.
A discussion, though short but important, with emphasis on the new features of the present work and
how to use the present work, is provided in section 4.

2. Derivation of force formula

In this section, we first state the flow field covered by the present study. Then we use a momentum
approach based on a suitably designed control volume to obtain a force formula (singularity approach).
Then this force formula is rewritten in a form such that the forces are only related to flow properties
inside the body and extended to multibody flow. Finally, an explicit force formula is obtained for
vortex production.

2.1. Description of the flow field

Consider a body with infinite span, immersed in an incompressible two-dimensional flow at at


constant density ρ . The freestream velocity V∞ is assumed horizontal. The local flow field is
supposed to be generated by vortices, sources, doublets and body acceleration and rotation, in a way
that the total velocity of the flow can be obtained by a linear superposition of the induced velocities due
to these factors. Singularities, including point vortices, sources and doublets, are assumed to be either
inside of the body (called inner ones) or outside of the body (called outer ones). The consideration of
vortices on the surface of the body due to vortex production will be pointed out specially when they are
considered. Each of the outer singularities will be assumed to be at a finite distance to the body.
The sum of the strengths of the inner vortices is equal to Γb = ∫
∂A
(udx + vdy ) , which is just the
circulation of the bound vortex, the closed curve ∂A is along the body with an anticlockwise path, so
that a clockwise circulation has a negative sign. We note that even when there is vortex production, the
conservation of total circulation holds
d Γi
∑i dt
= 0. (1)

The index i refers to any singularity independent of whether it is inside or outside of the body. In
section 2.2 we will show if (1) did not hold, then we would have a lift force whose magnitude depends
on the choice of reference frame.
A singularity, located at ( xi , yi ) but generally moving at the velocity ( dxi / dt , dyi / dt ) , will be
either a point vortex of strength Γi , a source of strength mi , or a doublet of strength μi .
The (fluid) velocity induced at ( x, y ) by a point vortex (i ) at ( xi , yi ) is
⎧ (i ) ∂ψ i ∂φi Γi y − yi
⎪u ( x, y ) = ∂y = ∂x = − 2π r 2
⎪ i
⎨ (2)
⎪ v (i ) ( x, y ) = − ∂ψ i = ∂φi = Γi x − xi
⎪⎩ ∂x ∂y 2π ri 2
Γi Γθ
where ψi = − ln ri with ri = ( x − xi ) 2 + ( y − yi ) 2 is the stream function and φi = i i
2π 2π
is the velocity potential. The angle θi is defined such that x − xi = ri cos θ i and
y − yi = ri sin θi . It should be emphasized that this induced velocity is independent of the velocity
of the vortex.

3
A doublet can be treated equivalently as a vortex pairError! Reference source not found.. Thus we first assume
the doublets have been transformed into vortices and just derive forces due to vortices and sources,
then the explicit influence due to doublets will be derived directly from the vortex based forces .
Pure source (sink) singularities are in fact not required since we only consider closed bodies, for
which we may always use a number of source doublets to represent pure sources. But for completeness
we will also consider the existence of sources. The flow field due to a point source (i ) of strength
mi is
⎧ (i ) ∂ψ i ∂φi mi x − xi
⎪ u ( x, y ) = ∂y = ∂x = 2π r 2
⎪ i
⎨ (3)
∂ψ ∂φ
⎪v ( i ) ( x, y ) = − i = i = i − yi
m y
⎪⎩ ∂x ∂y 2π ri 2
Since the functional forms for for (3) and (2) are similar, the forces due to sources can be similarly
obtained as for vortices.
Now consider body generated flow. For a body ( A ) rotating at the angular speed Ω (positive if
anticlockwise) around the point ( xo , yo ) which translates in addition at velocity (U , V ) in a flow
already with a free stream velocity V∞ . The flow potential and stream function due to body translation
and rotation may be decomposed as

φb = U φU + V φV + ΩφΩ , ψ b = Uψ U + Vψ V + Ωψ Ω (4)
where φU , φV , φΩ and ψ U ,ψ V ,ψ Ω are the so called normalized potentials and stream functions,
generated by the body translating and rotating at unitary speed. The forces due to this will be related to
added mass effects.

Fig. 1. Control volumes: the solid box ABCD defines a vertical control volume ( X 1 , X 2 ) × (−∞, ∞)
for lift study and the dashed box abcd defines a horizontal control volume (−∞, ∞) × (Y1 , Y2 )
for drag study.

2.2. Momentum approach for force analysis and force formula related to singularity velocity

Now we use a momentum balance approach, based on the control volumes (vertical control volume
and horizontal control volume) defined in Fig.1, to study the forces. The use of two types of control
4
volumes, for lift and drag, is to remove mathematical singularity manipulation in deriving force
formulas.
We assume the boundaries of the control volume to be far enough away from the body and
singularities, so that the momentum balance approach used here will be linear and therefore the
contributions by various singularities (when the body is regarded fixed) and by body generated flow
(when the body is accelerating and rotating) can be decomposed.
A) forces due to vortices. For lift, we use the vertical control volume for momentum balance. The
body is subjected to a lift force ( Lv ) , due to vortices, so that the fluid in the control volume is
subjected to a force of equal magnitude but with an opposite direction. This force is balanced by the
momentum flux across the left and right boundaries ( x = X 1 , x = X 2 ) and the time variation of the
momentum inside the control volume excluding the body, hence
dm(yi )
Lv = ρ ∑ ∫ V∞ v (i ) dy − ρ ∑ ∫ V∞ v (i ) dy − ∑ + Lav (5)
i
x = x1
i
x = x2
i dt
The last two terms on the right hand side represent the momentum change in the control volume
excluding the region occupied by the body. Hence if dm(yi ) / dt is defined for the whole space in the
control volume, i.e.,
X2 ∞
dm(yi ) d
∫ ∫ ρv
(i )
= dydx
dt dt X1 −∞

then, Lav represents the momentum variation rate of the fictitious fluid inside the body. The specific
role of Lav will be further discussed in the end of this subsection.
To find the explicit form of the integrals involved in (5), we use the identity
c
∞ πc

b + (y − d)
−∞ 22
dy =
b
which holds for any set of constants b, c, d independent of y . Hence
⎧ ∞ ρV∞ Γi X 1 − xi 1
⎪ ρ ∫ V∞ v X dy = = − ρV∞ Γi
(i )

⎪ −∞ 1 2 X 1 − xi 2
⎪ ∞ ρV Γ X − x 1
⎪ ρ ∫ V∞ v (i ) dy = ∞ i 2 i = ρV∞ Γi
⎪ −∞ X2 2 X 2 − xi 2
⎪ X
⎪d 2 ∞ 1 d⎛
xi
π ( x − xi )
X2
π ( x − xi ) ⎞

⎪ dt X∫ −∞
∫ 2π dt ⎜ X∫ x − xi
i ∫
ρ v (i )
dydx = ρ ⎜ Γ + Γ ⎟ dx
x − xi ⎟
i

⎪ 1 ⎝ 1 xi ⎠
⎪ 1 d
⎪ =ρ
2π dt
( −π ( xi − X 1 ) Γi + π ( X 2 − xi ) Γi )

⎪ d ( Γi xi ) 1 dΓ
⎪ = − ρΓi +ρ (π ( X 2 + X 1 ) ) i
⎩ dt 2π dt
Inserting these formulas into (5) we get
⎛ d ( Γi xi ) ⎞ 1 d Γi
Lv = − ρ ∑ ⎜ V∞ Γi − ⎟ + Lav + ( X 2 + X 1 ) ρ ∑
i ⎝ dt ⎠ 2 i dt
The last term on the right hand side has the factor X 2 + X 1 which is frame-dependent. If this
term did not vanish, we would have a force whose magnitude depends on the choice of the frame.
Thanks to (1) the last term on the right hand side vanishes so that the lift force is given by
5
⎛ d ( Γi xi ) ⎞
Lv = − ρ ∑ ⎜ V∞ Γi − ⎟ + Lav (6)
i ⎝ dt ⎠
Similarly, by using the horizontal control volume, we obtain the drag force formula

d ( Γi yi )
Dv = − ρ ∑ + Dav (7)
i dt
where Dav represents the change of x -momentum inside the body. Note that, if we had used the
vertical control volume for drag, then we may have singularity in manipulating integrals. This is the
reason to use two different control volumes for lift and drag.
B) forces due to sources. Comparing (3) to (2), it is obvious that the functional form of the velocity
components u and v due to a point source is the same as that of v and −u for a point vortex.
Hence we can use the results of vortices and directly write down the force formulas for point sources as
d ( mi yi ) d ( mi xi )
Lst = ρ ∑ + Las , Dst = ρ ∑ + Das (8)
i dt i dt
Here Las and Das represent the momentum change inside the body due to sources.
C) momentum change inside the body. Now consider the momentum changes inside the body due
to vortices ( Lav , Dav ) and sources ( Las , Das ). With La = Lav + Las and Da = Dav + Das we may
write
d d
La = ∑ ∫∫ ρ vdydx,Da = ∑ ∫∫ ρ udydx (9)
i dt A i dt
A

with the body fixed since the role due to accelerating translation and rotation will be treated separately
below. In appendix A, we will prove that
La = 0, Da = 0 (10)
D) body acceleration and rotation, added mass effect. The forces ( Ladd , Dadd ) due to body
acceleration and rotation have been well studied in the past using either the kinetic energy method (cf
LambError! Reference source not found.) or the unsteady Blasius equation (cf. Wu, Yang &Young27), and have
been shown, by Wu, Yang &Young27, as
⎧ d
⎪⎪ Ladd = − dt (UAuu + VAuv + ΩAuΩ )
⎨ (11)
⎪ D = − d (UA + VB + ΩB )
⎪⎩ add dt
uv vv vΩ

where Auu , Auv , AuΩ and Buu , Buv , BuΩ are added mass coefficients. The general method for
computing added mass coefficients can be found in LambError! Reference source not found. and Landweber &
ChwangError! Reference source not found..
E) Summary. Summing the force components defined in (6),(7), (8) and (10), we obtain the lift and
drag forces as
⎧ ⎛ d ( Γi xi ) ⎞ d ( mi yi )
⎪ L = − ρ ∑ ⎜ V∞ Γi − ⎟+ ρ∑ + Ladd
⎪ i ⎝ dt ⎠ i dt
⎨ (12)
⎪ d ( Γi yi ) d ( mi xi )


D = − ρ ∑i dt
+ ρ ∑ i dt
+ Dadd

6
Here the sum ∑ i
is performed over all the inner and outer singularities.

The force formulas defined by (12) (singularity velocity method) is rather general, including the
contribution from both the inner and outer singularities. The way to express the forces and the way to
obtain these formulas are new. In the case that Γ i is constant and the body is fixed, the force formulas
(12) can be simplified as
⎛ dx ⎞ dyi
L = − ∑ ρ ⎜ V∞ − i ⎟ Γi , D = − ∑ρ Γi (13)
i ⎝ dt ⎠ i dt
In the special case that the outside vortices, for instance the starting vortices, move at the freestream
speed V∞ and the internal vortices are fixed, as in the case of steady flow, we recover the classical
Kutta Joukowski theorem from (13)
L = − ρV∞ Γb,D = 0
Here Γ b = ∑Γ
in
i with ∑in
referring to summation over all the vortices inside the body.

2.3. Analysis of singularity interaction and effect of induced velocities

In (12), the forces are related to the speeds of the singularities. Now we first replace the speeds of the
free singularities by induced velocities and then perform analysis of the interaction between various
singularities to derive force formulas in terms of induced velocities so that we can do force
decomposition. For convenience, we use the condition (1) to make the term − ρ ∑ iV∞ Γi in (12)
disappear and split the right hand side of (12) into the following form

⎧ dx j dy
⎪L = ρ ∑ Γ j + ρ ∑ j m j + Lr
⎪ j , ou dt j , ou dt
⎨ (14)
⎪D = −ρ dy j dx
⎪ ∑
j , ou dt
Γ j + ρ ∑ j m j + Dr
j , ou dt

where
⎧ ⎛ d ( xk Γ k ) d ( mk yk ) ⎞ ⎛ dΓ j dm j ⎞
⎪ Lr = ρ ∑ ⎜ + ⎟ + ρ ∑⎜ xj + yj ⎟ + Ladd
⎪ k ,in ⎝ dt dt ⎠ j , ou ⎝ dt dt ⎠
⎨ (15)
⎪ ⎛ d ( yk Γ k ) d ( mk xk ) ⎞ ⎛ dΓ j dm j ⎞
⎪ D r = − ρ ∑ ⎜ − ⎟ − ρ ∑ ⎜ y j − x j ⎟ + Dadd
⎩ k , in ⎝ dt dt ⎠ j , ou ⎝ dt dt ⎠
The symbol ∑
j , ou
means summation over all the vortices and sources outside of and on the surface

of the body, while ∑


k ,in
means for those inside the body.

The velocity ( dx j / dt , dy j / dt ) for the motion of any free singularities involved in (14) is due to
free-stream convection and induction by all the inner and outer singularities except itself, i.e.,
⎧ dx j
⎪ = V∞ − ∑ ( Γ k Y jk − mk X jk ) − ∑ ( ΓlY jl − ml X jl )
⎪ dt k ,in l ,l ≠ j , ou

⎪ dy j =
⎪ dt ∑ (Γk X jk + mkY jk ) + ∑ (Γl X jl + mlY jl )
⎩ k ,in l ,l ≠ j , ou

7
Here X jk , Y jk are defined by
x j − xk y j − yk
X jk = , Y jk =
2π d 2
lj 2π d 2jk
with d 2jk = ( x j − xk ) 2 + ( y j − yk ) 2 . Inserting the expressions for (dx j / dt , dy j / dt ) into (14) ,

we may write the resulting force formulas as


⎧ L = Lb + Lif + L ff + Lt + L p + Ladd
⎨ (16)
⎩ D = Db + Dif + D ff + Dt + D p + Dadd
where the various components on the right hand sides have their own physical meaning, and are given
and discussed below.
a) The component ( Lb , Db ) is
Lb = ρV∞ ∑Γ j , Db = 0
j , ou

Since − ∑Γ
j , ou
j is equal to the total circulation Γ b of the bound vortices, we have Lb = − ρV∞ Γb

and Db = 0 . This is just the basic force (bound vortex force) given by the Kutta-Joukowski theorem.
b) The force ( Lif , Dif ) defined as

⎧ ⎛ ⎞
⎪ Lif = − ρ ∑ ⎜ ∑ ( Γ k Y jk − mk X jk ) Γ j − ∑ ( Γ k X jk + mk Y jk ) m j ⎟
⎪ j , ou ⎝ k ,in k ,in ⎠

⎪D = −ρ ⎛ ⎞
∑ ⎜ ∑ ( Γ k X jk + mk Y jk ) Γ j + ∑ ( Γ k Y jk − mk X jk ) m j ⎟
(f)
⎪ if j , ou ⎝ k ,in
⎩ k ,in ⎠
is due to the interaction between the inner singularities and outer singularities. Putting those terms
with the common factor Γ k (and similarly mk ) together and then exchanging the order of the double
sum, as
⎧ Lif = − ρ ∑Γ k ∑ (Y jk Γ j − X jk m j ) + ρ ∑mk ∑ ( X jk Γ j + Y jk m j )
⎪ k ,in j , ou k ,in j , ou

⎪ Dif = − ρ ∑Γ k ∑ ( X jk Γ j + Y jk m j ) − ρ ∑mk ∑ (Y jk Γ j − X jk m j )
(f)

⎩ k ,in j , ou k ,in j , ou

we obtain
Lif = − ρ ∑uk Γ k − ρ ∑mk vk , Dif = ρ ∑vk Γ k − ρ ∑uk mk (17)
k ,in k ,in k ,in k ,in

where (uk , vk ) , defined as


uk = ∑ (Y jk Γ j − X jk m j ) , vk = − ∑ ( X jk Γ j + Y jk m j ) (18)
j , ou j , ou

is the induced velocity at the location of the inner singularity ( k ) , that is, fluid velocity induced by all
the outside singularities.
c) The force ( L ff , D ff ), defined as

⎧ L ff = ρ ∑ ∑ ( ΓlY jl − ml X jl )( m j − Γ j )
⎪ j , ou l ,l ≠ j , ou

⎪ D ff = − ρ ∑ ∑ ( Γl X jl + mlY jl )( m j + Γ j )
⎩ j , ou l ,l ≠ j , ou

is due to the mutual interaction between the free singularities. It can be checked that the contributions

8
to this force by each pair of j , l with j ≠ l mutually cancel and thus
L ff = D ff = 0
Hence mutual interaction between the free singularities does not contribute to forces. Note that Wang
and Wu30 found this property for interaction between vortex rings when they study lift forces for
flapping flight.
d) The force component ( Lt , Dt ) defined as

⎛ d ( xi Γi ) d ( yk mk ) ⎞ ⎛ d ( yk Γ k ) d ( xk mk ) ⎞
Lt = ρ ∑ ⎜ − ⎟ , Dt = − ρ ∑ ⎜ + ⎟
k ,in ⎝ dt dt ⎠ k ,in ⎝ dt dt ⎠
is due to the motion and production of strengths of the inner singularities.
e) The force component ( L p , D p ) defined as

⎛ dΓ j dm j ⎞ ⎛ dΓ j dm j ⎞
Lp = ρ ∑ ⎜ x j − yj ⎟ , Dp = − ρ ∑ ⎜ y j + xj ⎟ (19)
j , ou ⎝ dt dt ⎠ j , ou ⎝ dt dt ⎠
is due to production of vortices and sources outside of the body. If there are no vortex production, then
d Γ (j f )
=0
dt
for each free vortex and L p = D p = 0 .

f) In the above derivation, doublets have not been considered and can be regarded as being
grouped into vortices, since each doublet can be represented by a vortex pair. Now we make the
force contribution due to possible doublets (inside the body) in an explicit form. As shown in
Thomson21, each doublet of strength μi and at position ( xi , yi ) can be considered as a vortex
pair of strength ∓ Γ i at ( xi , yi ± ε ) with ε → 0 and 2εΓi = μi . Apply (17) to the
corresponding vortex pairs yields, for ε → 0 , a force component ( Lu , Du ) of the following
form
∂ui( μ ) (μ )
I d ∂vi
Lμ = ρ ∑ i =1 2εΓi( μ ) , Dμ = − ρ ∑ i =1
I
d
2εΓi( μ )
∂y ∂y
or
∂ui ∂v
Lμ = ρ ∑ μi , Dμ = − ρ ∑ i μi
i , in ∂y i , in ∂y

The various force components above will be put in compact form in the next section. Before doing
this we would like to remark that the above analysis appears to have given a way to interpret the
physical origin of each force component. This is not seen elsewhere according to the knowledge of the
present authors. The force due to vortex production will be further discussed in section 2.5.

2.4. Summary of the force formulas based on induced velocity and multibody extension

Now assume there are vortices (not representing doublets now), sources and doublets inside the
body, and outside of the body there are a number of free vortices and sources. Inserting the force
components defined in items a)-f) of section 2.3 into (16), we obtain the final force formulas below,
as required by the paper,
⎧ L = − ρV∞ Γb + Lind + Lt + Lp + Ladd
⎨ (20)
⎩ D = Dind + Dt + D p + Dadd
9
Here ( Lind , Dind ), defined as
⎧ ∂ui
⎪ Lind = −∑ρ ui Γi − ∑ρ vi mi + ρ ∑ ∂y μi
⎪ i ,in i ,in i ,in
⎨ (21)
⎪ D = ρ v Γ − ρ u m − ρ ∂vi μ
⎪⎩ ind ∑ i ,in
i i ∑
i ,in
i i ∑
i ,in ∂y
i

is due to the induced velocity effect at the location of inner singularities, and (ui , vi ) , defined by (18)
and rewritten here as
⎛ Γ ( y − y ) m (x − x ) ⎞ ⎛ Γ (x − x ) m ( y − y ) ⎞
ui = ∑ ⎜ − j i 2 j + j i 2 j ⎟ , vi = ∑ ⎜ j i 2 j + j i 2 j ⎟ (22)

j , ou ⎝ 2π d ji 2π d ji ⎠⎟ ⎜
j , ou ⎝ 2π d ji 2π d ji ⎟⎠
denotes the fluid velocities induced at the location of the inner singularities by all the outside
singularities, not including induction by other inner singularities. The force component ( Lt , Dt ), due
to motion and production of inner singularities, is given by
⎛ d ( xi Γi ) d ( yi mi ) ⎞ ⎛ d ( yi Γi ) d ( xi mi ) ⎞
Lt = ρ ∑ ⎜ − ⎟ , Dt = − ρ ∑ ⎜ + ⎟ (23)
i ,in ⎝ dt dt ⎠ i ,in ⎝ dt dt ⎠
Finally the component ( L p , D p ), defined by (19), is due to production of singularities outside (and

on) of the body and will be further discussed in section 2.5, where we will show that only those
vortices in production on the surface of the body will contribute to forces.
Now we extend the force formulas (20) to the case of multiple bodies. Consider, for the case of
multiple bodies (namely body A , B , C ,  ), the force formulas for the body A with contours
∂A . Note that, the introduction of bodies B, C, … into the flow field around body A will induce a
perturbation (called body induced velocity) of the flow field in the same way as free singularities.
Moreover, the body induced velocity, due to any outside body, will be the same if this outside body be
replaced (or represented) by a number of equivalent singularities (bound and image sources and
vortices). These equivalent singularities are bounded to that body and cannot be regarded as free ones.
The force formulas (20) have been obtained for a single body without using pressure integration. If
the pressure p is used to integrate the force, as
L = ∫ pdx, D = − ∫ pdy
∂A ∂A

we of course should have the same forces as given by (20). Now remark that the forces in the form of
(20) only depend on the induced velocities inside the body and the motion and production of
singularities inside and on the body (the vorticity of a vortex conserves once it is in the flow, see
section 2.5) and that the flow pattern (such as induced velocity and their derivatives) inside the body is
the same no matter whether the outside singularities are free ones or equivalent ones (representing
outside bodies). We thus have the same force formula as given by (20) when there are outside bodies,
provided the flow induced by the outside bodies be regarded as induced by equivalent singularities
representing the outside bodies.
Thus, following (20) , the force formulas for body A can be written as
⎧⎪ LA = − ρV∞ Γb( A) + L(indA) + L(t A) + L(pA) + L(add
A)

⎨ (24)
⎪⎩ DA = Dind + Dt + D p + Dadd
( A) ( A) ( A) ( A)

where Γb( A) is the circulation around body A ,

10
⎧ ( A) ∂ui
⎪ Lind = −∑ρ ui Γi − ∑ρ vi mi + ρ ∑ ∂y μi
⎪ i, A i, A i, A

⎪ D ( A) = ρ v Γ − ρ u m + ρ ∂ui μ
⎪⎩ ind ∑ i, A
i i ∑
i, A
i i ∑
i , A ∂x
i

is the induced velocity effect, with summation performed over all singularities inside body A , and
(ui , vi ) is the velocity at ( xi , yi ) (position of singularities inside body A) induced by all the outside
( A) ( A)
singularities and bodies. The unsteady term ( Lt , Dt ), now defined by
⎛ d ( xi Γi ) d ( yi mi ) ⎞ ( A) ⎛ d ( yi Γi ) d ( xi mi ) ⎞
L(t A) = ρ ∑ ⎜ − ⎟ , Dt = − ρ ∑ ⎜ + ⎟
i, A ⎝ dt dt ⎠ i, A ⎝ dt dt ⎠
is due to the motion and production of singularities inside body A . Finally, the force component
( A) ( A)
( L p , D p ) is due to the vortex production (at for instance geometric singularities) on the surface of

body A and will be further discussed in section 2.5.


The force formulas for bodies B , C ,  can be similarly defined.
Remark Compared to Wu, Yang &Young27, where the force formula has been obtained directly
through the unsteady Blasius equation (suitable only for irrotational flow so that it does not apply to the
case when vortices are produced on the surface of the body), the force formulas (20) and (24) based on
the induced velocity method are more general since here we have included the role of bound vortices
and vortex production. Moreover, the induced velocity in the force formula of Wu, Yang &Young27 is
due to all the singularities (including the inner ones) while here this induced velocity is only due to
outside singularities (and bodies). The multibody force formulas (24) will be checked using the two
11
cylinder example of Crowdy in the next section.

2.5. Force due to vortex production

Now we consider the force component ( L p , D p ), defined by (19) and due to production of

singularities outside of the body. Here we derive an explicit form for this force, which is frame
independent and convenient for identifying the role of vortex production.
Due to the Kelvin theorem of conservation of circulation, we have d Γ j / dt = 0 once a vortex (j)

is produced and is moving freely outside of the body. Hence it remains only those just in production on
the surface of the body. Generally, vortices will be produced at some geometric singularities, such as
the trailing edge of a body. Here we do not consider source production outside of the body. Then by
(19), we may write
d Γs d Γs
Lp = ρ ∑ xs , D p = − ρ ∑ ys (25)
s dt s dt
Here the summation is performed over the points on the surface of the body where we have vortex
production at rate d Γ s / dt .
( A) ( A)
For multibody problems, the force components ( L p , D p ) due to the vortex production on the

surface of body A are computed as


d Γs d Γs
L(pA) = ρ ∑ xs , D p( A) = − ρ ∑ ys (26)
s, A dt s, A dt
The force due to vortex formation, in the form of (25), is strange due to the dependence on the
position xs and ys . This would mean that the magnitude of the forces depend on the choice of the

11
reference frame. This is in fact not so since in real problems the vortices are always produced in pair
due to conservation of vorticity. This means that the production of one vortex of circulation Γ n at
some location ( xn , yn ) (which may be some point near the trailing edge of an airfoil) is at the

consequence of the production of another vortex of circulation (


−Γ n at a point x n , y n ) (which

may be a point inside the body close to the trailing edge) close to ( xn , yn ) . The y momentum due
to this vortex pair is
X2 ∞
⎛Γ Γ
m y (t ) = ∑
n
∫ ∫ ρv
X1 −∞
(n)

k ⎝ 2 2
( ⎞
dydx = ∑ρ ⎜ n ( X 1 + X 2 − 2 xn ) − n X 1 + X 2 − 2 x n ⎟

)
or
m y (t ) = ∑ρΓ n ( x n − xn )
n
Hence the force due to this production is
d Γn  d Γn 
Lp = −∑ρ
n dt
( x n − xn ), D p = ∑ρ
n dt
(
y n − yn ) (27)

Hence the magnitude of the force due to vortex production is frame independent. The way to treat
vortex production outside the body in the way of expression (25) or (27) is very convenient. In the next
section we will use the example of Karman vortex street to demonstrate the usefulness of (27). Perhaps
the writing of the force due to production in the form (26) is the only possible way to have force
decomposition for multibody flow with vortex production.

3. Validation study

The application of the force formulas given in section 2 relies on the knowledge of strength, location
and speed of singularities, including equivalent singularities representing outside bodies. Once these
flow structures are known, the force formulas presented in section 2 can be used to obtain the forces or
identify the role of each flow structure. In this section we give several examples, for which the flow
structures are assumed or can be obtained analytically, to demonstrate the application of the force
formulas. In section 3.1 we apply the induced velocity method to circular cylinders for validating the
induce velocity method (20), and notably the multibody force formulas (24). For circular cylinders
distribution of singularities can be determined by the method of images. In section 3.2 we consider the
drag for Karman vortex street, which involves vortex production so that we may use it to check the
force formula (27). This problem involves in addition an infinite number of discrete vortices with
assumed pattern independent of the shape of the body. Hence for the part of drag other than the vortex
production drag, we can not apply the induced velocity formulas (20). Instead, we have to use the
singularity velocity formulas (12).

3.1. Problems of circular cylinder

Here we first simplify the induced velocity method for the case of circular cylinders, then we apply
the results to study one cylinder with a pair of outside standing vortices and a source doublet, and the
problem of two circular cylinders with given circulation. All the problems have known solutions so that
they are used to validate the present force formulas.

3.1.1. Simplified force formula for a circular cylinder


12
For each vortex j of circulation Γ (j f ) outside of a circular cylinder, there is one image vortex of
circulation Γ (jo ) = Γ (j f ) at the origin, and one of circulation Γ (jm ) = −Γ (j f ) at the inverse point
( x (jm ) , y (jm ) ) . An outside source doublet at ( x (jσ ) , y (jσ ) ) and with strength σ j , has an image doublet
of strength
a2
σ (j m ) = σj
f j2
at its inverse point, where f j is the distance of the outside doublet to the body center. The force

formula (20) applied here yields the following force decomposition


⎧ L = LB + Lb + Lo + Lm + Lμ + L f + Ln

⎩ D = DB + Db + Do + Dm + Dμ + D f + Dn
with
⎧ LB = − ρV∞ Γb ,DB = 0
⎪  (b )  (b )
⎪ Lb = − ρ ∑u i Γi ,Db = ρ ∑vi Γi
⎪ i ,in i ,in

⎪ 
J
⎪ oL = − ρ ∑ u i Γ (o)
i ,D o = ρ ∑ v i Γi( o )
⎪ i ,in l =1

⎪ Lm = − ρ ∑u l Γl ,Dm = ρ ∑v l Γl
 (m) (m)

⎨ i ,in i ,in (28)



⎪ Lμ = ρ ∑ ∂ui μi , Dμ = − ρ ∑ ∂vi μi
⎪ i ,in ∂y i ,in ∂y

⎪ L = ρ ∂u j a σ j , D = − ρ ∂v j a σ j
2 2

⎪ σ ∑j , ou ∂y f j2
σ ∑
j , ou ∂y f j2

⎪⎩ L f = 0, D f = 0
Here ( LB , DB ) is the basic bound vortex force, the components ( Lb , Db ), ( Lo , Do ) and
( Lm , Dm ) are due to induced velocities at the locations of bound vortices, image vortices at body
center and image vortices at inverse point, respectively. The component ( Lμ , D ) is due to the
μ

induced velocity gradient at the location of inner real doublets, and the component ( Lσ , Dσ ) is due to

the images (at the inverse point) of the outside doublets. The velocity (u , v ) is the induced fluid
i i

velocity (induced by all the outside vortices and source doublets) relative to the speed of the internal
d Γn 
singularity ( i ). Finally the component ( Ln , Dn ) with Ln = −∑ nρ ( x n − xn ),
dt
d Γn 
Dn = ∑ nρ
dt
(
y n − yn )  n , y ) denote the
is due to vortex production, ( xn , yn ) and ( x n

position of the vortex pair with circulation production.

3.1.2. Standing vortex pair behind a circular cylinder

It is well known that at moderate Reynolds numbers, the flow around a circular cylinder involves
two standing, oppositely rotating vortices behind its cylinder. An inviscid model for this consists of two
equal and opposite point vortices, of circulation Γ > 0 and −Γ < 0 , standing symmetrically behind
13
 
the cylinder12,21, at the positions x + = (rf cos θ f , rf sin θ f ) and x − = (rf cos θ f , − rf sin θ f ) ,
respectively.
 (m) a2 a2
The image vortices at the inverse points are respectively at x + =( cos θ f , sin θ f ) and
rf rf
 (m) a2 a2
x − = ( cos θ f , − sin θ f ) . On the Foppl line (see for instance Saffman12) defined by
rf rf

(r − a 2 ) = 4rf4 sin 2 (θ f ) or4rf4 − ( rf2 − a 2 ) = 4rf4 cos 2 (θ f ) , the two vortices, though under
2 2 2
f

the convection by stream flow and under induction by the vortices (including images) and source
doublet of strength μ = 2π a 2V∞ , remain stationary if the circulation is given by
⎛ a4 ⎞
Γ ± = ∓ Γ, Γ = 4π V∞ rf sin (θ f ) ⎜1 − 4 ⎟
⎜ r ⎟
⎝ f ⎠

It is well known that for this case the drag vanishes, either by direct calculation of pressure on the
cylinder or by considering the vortex pair as a source doublet at far enough distance12. Now we will
check if we recover this conclusion by the force formula in terms of the induced velocity. The induced
(m) (m) (o)
velocities v+ and v− at the two inverse points and the induced velocity v and its
∂v(o)
derivative at the center of the cylinder are found to be
∂y
⎛ a 2V∞ sin ( 2θ f ) ( m ) a 2V∞ sin ( 2θ f )
⎜ v+( m ) = , v− = −
⎜ rf2 rf2

⎜ v ( o ) = 0, ∂v = − Γ sin 2θ f
(o)

⎜ ∂y π rf2

The drag force due to the induced velocities at the two image points is, due to the force relation in
(28),
2 ρΓa 2V∞ sin ( 2θ f )
Dm = ρΓ (+m ) v+( m ) + ρΓ (−m ) v−( m ) = 2
r f

and the drag due to the doublet at the body center is, according to the fifth relation in (28),
∂v ( o ) (0, 0) 2a 2 ρV∞ Γ sin 2θ f
Dμ = − ρμ =−
∂y rf2
Hence Dm cancels Dμ , and the total drag obtained by the present method vanishes.

3.1.3 A doublet outside of a circular cylinder

Consider a doublet of strength σ at y = f , then the velocity induced by this doublet on the line
σ
x = 0 is u ( μ ) (0, y ) = so that its derivatives at the center and inverse point are
2π ( f − y ) 2
∂u ( μ ) σ ∂u (σ ) σ
(0, 0) = , (0, ym ) =
∂y πf ∂y π ( f − a2 / f )
3 3

According to the fifth and sixth relations in (28), if there is a doublet of strength μ at the center of
the cylinder, there is a lift force given by
14
∂u ( μ ) σμ
Lμ = ρ (0, 0) μ = ρ
∂y π f3
and there is an outside doublet, there is a lift due to the image of this doublet
∂u (σ ) σ a2 ρ a2 f
Lσ = ρ (0, ym ) 2 = σ2
∂y π ( f −a )
3
f 2 2

The latter is the same as that given by the Blasius theorem or by Lagally theorem (cf. Milne-
21
Thomson ). This is a force which points to the doublet.

3.1.4 Two circular cylinders with circulation

Crowdy11 gives a general theory to find the flow potential for the calculation of lift for a finite
number of staggered cylinders with bound vortex. Note that Crowdy did not give force formulas for
each individual cylinder. The forces are obtained using the Blasius equation once the flow potential is
obtained.
Here we apply (28) to obtain the lift for his example of two vertically aligned cylinders (see Fig.2),
both of radius a = 1/ 2 and of given circulation Γ1 = Γ 2 = Γ ( Γ = −5, −2, 0 ), immersed in a
uniform stream V∞ = 1 and placed at a distance h ≥ 1 between their centers. The bound vortex
associated with one cylinder has an image pair of two counter rotating vortices in the other, at the
center and inverse point respectively. An image pair in one cylinder also has two counter rotating image
vortices in the corresponding inverse points of the other cylinder, plus two cancelling images at the
center of the other cylinder cancel. Thus in each cylinder, there are an infinite number of such inverse
points, with counter rotating vortices of equal strength between any two adjacent inverse points, with
distance becoming closer for newer generated images. If, for the i th cylinder, the distance of the
k th inverse point to the center of this cylinder is denoted as hk(i ) , then
⎛ a2 ⎞
h1(i ) h = a 2 ,andhk(i ) ⎜ h − (i ) ⎟
= a 2 for k > 1 (29)
⎝ h − hk −1 ⎠

Fig. 2. Staggered two cylinder problem of Crowdy (2006). Left: staggered cylinders with bound vortices. Right:
vortex system with image vortices for the equivalent inverse point model.

For an exact solution with (28), we have to work with such infinite number of inverse points. Here
we instead use an approximate method based on the remark that the inverse points in each cylinder are

15
distributed in a narrow region. To see this, let hk
(i )
= g (i ) for k → ∞ , then by (29) we have
⎛ a2 ⎞
g (i ) ⎜ h − (i ) ⎟
= a 2 , which can be solved to give
⎝ h − g ⎠
1
(
g ( i ) = h − h 2 − 4a 2
2
) (30)
2 (i ) (i )
and one can verify that a / h < hk < g for k > 1 . It can be further verified that the region
between y = a 2 / h and y = g (i ) is very narrow even for h close to 1 .
Hence within the framework of approximate solution we may merge all the inverse points into an
(i )
equivalent one, with a distance he to the center of the corresponding cylinder satisfying
a / h < h < g . Moreover, since these inverse points are denser close to y = g (i ) , we just set
2 (i )
e
(i )

1
he(i ) = a 2 / h . With a = , we have he(i ) = he = 0.52 / h . Let hee be the distance between
2
the equivalent inverse points of the two cylinders, and heo be the distance between the equivalent
inverse point of one cylinder to the center of the other. Then
hee = h − 2he , heo = h − he
With the approximation of equivalent inverse point, and considering the conservation of circulation,
the vortex system in each cylinder can be simplified in the following way.
For the 1 st cylinder, there is one given bound vortex of circulation Γ1 = Γ1 and one image
(b )

vortex of circulation Γ1 = Γ 2 at the center of this cylinder, plus one equivalent image vortex of
(o)

circulation Γ = −Γ 2 at the equivalent inverse point y = he(i ) . For the 2nd cylinder, this can
(m)
1

be similarly defined. Moreover, each cylinder has a doublet of strength


π
μ1 = μ2 = μ = 2π a 2V∞ = . There are also an infinite number of image doublets at the inverse
2
points. When the equivalent inverse point is applied to the doublet, the strength of the doublet at this
inverse point is μ1( e ) = μ2( e ) = μ ( e ) ≈ μ a 2 / heo2 .
The velocities induced at the center and at the equivalent inverse points of the 1st and 2nd
cylinders by the bound and image vortices and doublet of the 2nd and 1st cylinders are thus
⎧ (o) 2Γ −Γ μ μ (e)
⎪ 1u = −u (o)
= + + +
2π h 2π heo 2π h 2 2π heo2
2


⎪u ( m ) = −u ( m ) = 2Γ + −Γ + μ + μ
(e)

⎪⎩ 1 2
2π heo 2π hee 2π heo2 2π hee2
The derivatives of the corresponding induced velocities are
⎧ ∂u1( μ ) ∂u2( μ ) 2Γ −Γ μ μ (e)
⎪ ∂y = − = + + +
⎪ ∂y 2π h 2 2π heo2 π h3 π heo3
⎨ (σ ) (σ )
⎪ ∂u1 = − ∂u2 = 2Γ + −Γ + μ + μ
(e)

⎪⎩ ∂y ∂y 2π heo2 2π hee2 π heo3 π hee3


According to (28), the force formula for the lower airfoil is L = LB + Lb + Lo + Lm + Lμ
(1) (1) (1) (1) (1) (1)

B = − ρV∞ Γ1 ,
L(1) b = − ρ u1 Γ1 ,
L(1) o = − ρ u1 Γ 2 m = − ρ u1 Γ 2
(o)
with L(1) (o) (o)
, L(1) (m) (m)
,

16
∂u1( μ ) ∂u1(σ ) ( e )
μ = ρ
L(1) μ1 , L(1)
σ = ρ μ .
∂y ∂y
1
With ρ = V∞ = 1 , Γ1 = Γ (2o ) = −Γ (2m ) = Γ and μ1 = μ = π
, we have
2
∂u ( μ ) ∂u (σ )
L(1) = −Γ − 2u1( o ) Γ + u1( m ) Γ + 1 μ + 1 μ ( e )
∂y ∂y
(μ ) (σ )
where the induced velocities u1 , u1 and their derivatives ∂u1 / ∂y, ∂u1 / ∂y have been
(o) (o)

given above. Similarly for the upper airfoil,


∂u2( μ ) ∂u (σ )
L(2) = −Γ − 2u2( o ) Γ + u2( m ) Γ + μ + 2 μ (e)
∂y ∂y
The results, compared to Crowdy11, are displayed in Fig 3. We remark that the agreement is
acceptable even when h is short and despite the use of equivalent inverse point to merge all the
inverse points. The short distance behavior, that is, there is an attraction force when the cylinders are
close, has been discussed by Crowdy. Here it is found that this is due to the influence of the real and
image doublets and the induced velocity gradient, described by Lμ and Lσ in (28).

Fig. 3. Comparison between the present result (solid lines) with Crowdy (2006, Fig.4)(dashed lines), for Γ=0
(lower), Γ = −2 (middle) and Γ = −5 (upper). Dashed lines: Crowdy.

3.2. Karman vortex street

The problem of the Karman vortex street behind a bluff body (Fig.4) is rather special since the shape
of the body is unknown. The Karman vortex street is a double row of staggered and counter-rotating
vortices of strength Γ > 0 and Γ < 0 and moving horizontally at speed
Γπ πb
V= tanh
a a
Here b is the vertical separation distance between these two rows and a is the horizontal separation
distance between adjacent two vortices in each row (see for instance Milne-Thomson21). The period of
vortex shedding in each row is thus

17
a
τ=
V∞ − V
Now consider the vortex production force. According to (27), the drag, averaged over τ , due to the
shedding of new vortex pair, separated at a distance y n − yn = b , is

( )
d Γ n y n − yn dt ρ
∫ d (Γ ( y )) = ρτbΓ
1 τ τ

τ∫
Du = ρ = − yn
0 dt τ 0 n n

When the expressions for V and τ are used, we obtain


ρ bΓ (V∞ − V ) ρ bΓ ⎛ Γπ πb ⎞
Du = = ⎜ V∞ − tanh ⎟
a a ⎝ a a ⎠
This is the well-known formula for the unsteady part of the drag, which has been otherwise obtained
by sophisticate Blasius approach21, impulse approach12 or integral approach20. The way to obtain this
force using the present approach (27) appears to be much easier. This allows us to validate the vortex
production force formula.

Fig. 4. Control volume for the Karman Vortex street problem. A cutline x = xd separates the flow regime into
an upstream part and a downstream part (D).

Remark also that, apart from the drag component Du due to shedding of new vortices, there is
also another component Ds due to quasi steady flow formed by two rows of periodically
counter-rotating vortices spreading infinitely into the downstream direction. No existing force formula,
including the present formula (12), can be directly applied to this case. One then needs to construct a
special downstream boundary and relates this component of drag to the momentum flux across this
boundary. For details, see Milne-Thomosn21, Saffman12 and Howe20, where different ways for setting
the downstream boundary are used. Let x = xd be such a boundary downstream of the body and
intersecting the vortex street in its uniform region. Let (uk , vk ) be the induced velocity by the vortex
street. Then, in Appendix B, we shall prove that, the use of (12) leads to

Ds =
1
2 ∫x = xd
ρ (uk2 − vk2 ) dy (31)

which is exactly the same as obtained by Howe20 through an integral approach. Howe then shows that
(31) yields

18
Γ2 ⎛ a πb ⎞
Ds = 2 ⎜
− b tanh ⎟
2a ⎝ π a ⎠
The total drag D = Du + Ds is thus
ρ bΓ (V∞ − V ) Γ2 ⎛ a πb ⎞
D= + 2 ⎜
− b tanh ⎟
a 2a ⎝ π a ⎠

4. Conclusions and discussions

Started from a momentum balance analysis based on suitably designed control volumes, which are
different for lift and drag to remove mathematical difficulties, and proceeded with interhange between
the singularity velocity and induced flow velocity, we have obtained force formulas for single and
multibody flows with multiple free and bound vortices and vortex production. The influence on the
force by vortex production is treated in a simple, explicit and frame-independent way, suitable for
identifying the role of vortex production (section 2.5).
The present work is new for three reasons. First, it covers the work of Wu, Yang &Young27 as a
special case, and includes in addition the effect of bound vortices and vortex production. Second, the
way to obtain the force formulas is based on the interaction of various singularities, and thus is useful
for explicitly interpreting the influence due to various resources. For instance, we have shown that the
interaction between free singularities do not contribute to forces, while the induced velocity effect is
due to interaction between free singularities and inner singularities (see section 2.3). Most importantly,
the force formulas are expressed in a physically arranged and explicit form which allow for
identification of roles of different contributions and force decomposition.
The force formulas in the form of (20) (single body) or (24) (multibody) can be used in several
ways.
First, we may use them to identify or understand the role of outside vortices/bodies and vortex
production on the forces of the actual body, since each contribution is made explicit and the various
terms in the formulas are arranged according to their physical origin. For instance, according to
whether the induced velocity due to an outside source increases or cancels the local velocity on the
actual body, we may identify that the lift force is increased or reduced, depending on the distance
between the outside source and the body surface. For instance, for the Crowdy example of two circular
cylinders, using the present induced velocity approach (24) it is found that the attraction force when the
two cylinders are close is due to the influence of the real and image doublets inside each cylinder and
the induced velocity gradient due to another cylinder.
Second, we may use the formulas (20) or (24) to optimize arrangement of outside vortices and bodies
for force enhancement or reduction, if a joint solver can be built for finding the flow structure (notably
singularity distribution). Optimization algorithm can always be adequately designed since the relation
between the forces and singularity properties is explicit in (20) or (24).
Finally, we may use the formulas to derive analytical force formulas once the flow field is given or
known. In the examples given in section 3 we do obtain analytical force formulas with (20), (24)
and (27), though some were known from other theories.
The force formulas (12), (20) or (24) may be put into integral forms through the use of Dirac
function, thus applicable to problems where the singularities are continuously distributed. This, along
with further applications including derivation of force formulas for particular problems, will be

19
considered in forthcoming works.

Acknowledgement

This work was supported by National Basic Research Program of China (2012CB720205).

References

1. Batchelor F.R.S. An introduction to fluid dynamics, Cambridge University Press,Cambridge;1967.


2. Crighton D.G. The Kutta condition in unsteady flow. Annual Review of Fluid Mechanics 1985;17:
411-445.
3. Anderson J. Fundamentals of Aerodynamics, Mcgraw-Hill Series in Aeronautical and Aerospace
Engineering, McGraw-Hill Education,New York; 2010.
4. Chow C.Y. & Huang M.K. The initial lift and drag of an impulsively started aerofoil of finite
thickness. Journal of Fluid Mechanics 1982;118:393-409.
5. Lee F.J. & Smith C.A. Effect of vortex core distortion on blade-vortex interaction. AIAA Journal
1991;29:1355-1362.
6. Aref H. Point vortex dynamics: a classical mathematics playground. Journal of Mathematical
Physics 2007;48:065401.
7. Oterberg D. Multi-body unsteady aerodynamics in 2D applied to a vertical-axis wind turbine using
a vortex method. Master Thesis, Uppsala Universtity, Uppsala: 2010.
8. Smith F.T. & Timoshin S.N. Planar flows past thin multi-blade configurations. Journal of Fluid
Mechanics 1996;324:355-377.
9. Katz J. & Plotkin A. Low Speed Aerodynamics. Cambridge University Press, Cambridge; 2001.
10. Hsieh C.T., Kung C.F.&Chang C.C. Unsteady aerodynamics of dragonfly using a simple
wing-wing model from the perspective of a force decomposition. Journal of Fluid Mechanics
2010;663:233-252.
11. Crowdy D. Calculating the lift on a finite stack of cylindrical aerofoils. Proceeding of the Royal
Society A. 2006;462:1387-1407.
12. Saffman P.G. Vortex dynamics. Cambridge University Press, New York; 1992.
13. K. Streitlien, & M.S. Triantafyllou. Force and Moment on a Joukowski Profile in the. Presence of
Point Vortices. AIAA Journal 1995;33 (4):603-610.
14. Ramodanov, SM. Motion of a circular cylinder and N point vortices in a perfect fluid, Regular and
Chaotic Dynamics 2002;7:291-298.
15. Shashikanth B.N., Marsden J.E., Burdick J.W. &Kelly S.D. The Hamiltonian structure of a
two-dimensional rigid circular cylinder interacting dynamically with N point vortices. Physics of
Fluids 2002;14:1214-1227.
16. Borisov, Alexey V.; Mamaev, Ivan S.; Ramodanov, Sergey M. Dynamic interaction of point
vortices and a two-dimensional cylinder, Journal of Mathematical Physics. JOURNAL OF
MATHEMATICAL PHYSICS 2007;48(6):065403.
17. Kanso E. & Oskouei B.G. Stability of a coupled body--vortex system. Journal of Fluid Mechanics
2008;600:77-94.
18. S. Michelin & S. G. Llewellyn Smith. Falling cards and flapping flags: understanding fluid-solid
interactions using an unsteady point vortex model. Theoretical and Computational Fluid Dynamics
2010;24:195–200.
19. Wu J.C. Theory for aerodynamic force and moment in viscous flows, AIAA Journal
1981;19:432-441.
20. Howe M.S. On the force and moment on a body in an incompressible fluid, with application to
rigid bodies and bubbles at high Reynolds numbers, Quartly. Journal of Mechanics and Applied
Mathematics 1995;48:401-425.
21. Milne-Thomson L.M. Theoretical Hydrodynamics, Macmillan Education LTD, Hong Kong; 1968.
22. Ragazzo, C. G. & Tabak, E. G. On the force and torque on systems of rigid bodies: a remark on an
integral formula due to Howe. Physics of Fluids 2007;19:057108.
23. Chang C.C., Yang S.H. & Chu C.C. A many-body force decomposition with applications to flow
about bluff bodies. Journal of Fluid Mechanics 2008;600:95-104.
24. Sakajo T. Force-enhancing vortex equilibria for two parallel plates in uniform flow, Proceedings of
the Royal Society A 2012;468:1175–1195.
25. Bai CY & Wu ZN. Generalized Kutta-Joukowski Theorem for multi-vortices and multi-airfoil flow
(lumped vortex model). Chinese Journal of Aeronautics;2014 27(1): 34–39
20
26. Landweber L & Miloh T. Unsteady Lagally theorem for multipoles and deformable bodies, Journal
of Fluid Mechanics 1980;96:33-46.
27. Wu C.T., Yang F.L. & Young D.L. Generalized two-dimensional Lagally theorem with free
vortices and its application to fluid-body interaction problems. Journal of Fluid Mechanics
2012;698:73-92.
28. Lamb H. Hydrodynamics, Dover Publications, New York; 1932.
29. Landweber L & Chwang A. Generalization of Taylor's added-mass formula for two bodies. Journal
of Ship Research 1989;33:1-9.
30. Wang X.X. & Wu Z.N. Stroke-averaged lift forces due to vortex rings and their mutual interactions
for a flapping flight model. Journal of Fluid Mechanics 2010;654:453-472.
31. Van der Geer J, Hanraads JAJ, Lupton RA. The art of writing a scientific article. J Sci Commun
2000;163(1):51–9.

BAI ChenYuan is a PhD student at the School of Aerospace Engineering, Tsinghua University.
E-mail: baichenyuan@sina.com

LI Juan is a PhD student at the School of Aerospace Engineering, Tsinghua University.


E-mail: li-juan13@mails.tsinghua.edu.cn

WU ZiNiu is a professor and Ph.D. supervisor at the department of engineering mechanics, Tsinghua University,.
Her current research interests are aerodynamics, notably vortex flow and high speed flow with shock reflection and
interaction.
E-mail: ziniuwu@tsinghua.edu.cn

Appendix A Momentum change inside a fixed body

Now will show that

d d
La 
dt ∫ ∫A
ρvdydx = 0, Da  ∫ ∫ ρudydx = 0
dt A
(32)

when the body A is considered fixed.

For each vortex or source i inside the body, we define an infinitesimal fixed circle of
radius rt  ( x − xi ) 2 + ( y − yi ) 2 % = ε whose center instantaneously coincides with the vortex.
For convenience we use here the subscript x, y and t denote the partial differentials with respect
to x ,y and t . Let Σ be the region bounded by the contour of the body ∂A and the
circumferences of the circles ri = ε . Below are some additional relations.

Decompose La defined in (32) as

d d
La = ρ ∫ ∫ vdydx+ρ ∑ ∫ ∫ vdydx
dt Σ ri <ε
i,in dt

Since v = −ψ x and since ψ is analytical in Σ , we use the divergence theorem to write

∫ ∫ vdydx = − ∫ ∫ ψ x dydx = − ∫ ψ dy + ∑ ∫ ψ dy
Σ Σ
∂A i ,in ri =ε

Hence

d d d
La = − 
∫ ψ dy + ∑ % ∫ ψ dy + ρ ∑ ∫ ∫ vdydx
dt ∂A ri <ε
i ,in dt ri =ε i ,in dt

21
Consider the last term on the right hand side. Across an element of length ri dθi at ( ri , θ i ) on the
circumference of the circle ri < ε , the loss of momentum due to the movement of the vortex (i ) is

dmi = −v (i ) ( xi ,t cos θi + yi ,t sin θi ) ri dθi

Γi cos θi
Remark that v
(i )
= and
2π ri

d 2π Γ 2π
∫ ∫ vdydx = ∑ ∫ dmi = −∑ i ∫ cosθi ( xi ,t cos θi + yi ,t sin θi ) dθi
dt ri <ε i ,in 2π
0 0
i ,in

Hence

d Γi
∑ dt ∫ ∫
i ,in
r ( i ) <ε
vdydx = −∑
i ,in 2
xi ,t

If sources can be similarly analyzed. When both vortices and sources are present we may write

d ⎛Γ m ⎞
∑ dt ∫ ∫
i ,in
r ( i ) <ε
vdydx = −∑ ⎜ i xi ,t − i yi ,t ⎟
i ,in ⎝ 2 2 ⎠

Now consider the second term on the right hand side. First consider vortices.

∂ ⎛ Γi ⎞ Γ
ψ t(i ) = − ⎜ ln ri ⎟ = − i ri ,t
∂t ⎝ 2π ⎠ 2π ri

Let ( x, y ) be a fixed point on the circle. Differentiating x − xi = ri cos θi and


y − yi = ri sin θi with respect to time we obtain

riθi ,t = xi ,t sin θi − yi ,t cos θi , ri,t =-x i,t cos θi -yi,t sin θi

2π 2π
Remark also that ∫ 0
cos 2θ (i ) dθ ( i ) = π and ∫ cosθ (i ) sin θ (i ) dθ (i ) = 0 . Thus and similarly
0
we have

Γi 2π Γi
∫ε ψ
ri =
i ,t dy = −
2π ∫ 0
ri ,t cos θi dθi =
2
xi ,t

Hence

d Γi
∑ dt ∫ ψ dy = ∑ 2 xi ,t (33)
i ,in ri =ε i ,in

When sources are included, the analysis is similar and we may write

d ⎛ Γi mi ⎞
∑ dt ∫ ψ dy = ∑ ⎜⎝ 2 xi ,t −
2
yi ,t ⎟

i ,in ri =ε i ,in

22
In summary we have proved

d
dt 
∫ ψ dy
La = −
∂A

Since here the body is assumed stationary so that ψ is a constant along the body, and
thus ∫ ψ dy = 0 . This means
∂A
La = 0

Similarly we may prove Da = 0

Appendix B Additional expressions for Karman vortex street

To prove (31) using the force formula (12), we need a relation between the velocities of the inner and
outer vortices. For this purpose will define a large contour ∂C enclosing the body and a part of outer
vortices. Moreover, for each outer vortex inside ∂C , we define an infinitesimal fixed circle of radius
ri = ε whose center instantaneously coincides with the vortex. Now consider the fluid region 
enclosed by ∂C , ∂A and Λ , where Λ denote the perimeters of all the fixed circles ri = ε . We
derive some integrals along the contours ∂ .

Since φ and ψ are analytical in  , we may use the divergence theorem and the identity
φ y ≡ −ψ x , φx ≡ ψ y to write

∫ (φ dx −ψ dy ) = − ∫ ∫ (φ y +ψ x )dxdy = 0

∂

∫ (φ dy +ψ dx ) = ∫ ∫ (φ
∂
 x −ψ y )dxdy = 0

Hence

⎧ (ψ dy − φ dx ) − (φ dx −ψ dy ) = (ψ dy − φ dx )
⎪ ∫
∂A
Λ∫ ∫
∂C
⎨ (34)
⎪ ∫ (ψ dx + φ dy ) + ∫ (φ dy + ψ dx ) = ∫ (φ dy +ψ dx )
⎩ ∂A Λ ∂C

ri = ε
Similarly, for each vortex inside ∂A , we define an infinitesimal fixed circle of radius
whose center instantaneously coincides with the vortex. Now consider the region Σ enclosed by ∂A
and Θ , where Θ denote the perimeters of all the fixed circles ri = ε inside ∂A . Since φ
and ψ are analytical inside the region Σ , we can apply the divergence theorem to

∫ (ψ dy − φ dx )
∂A
to write

∫ (ψ dy − φ dx ) − ∫ (ψ dy − φ dx ) = ∫∫ (φ
∂A Θ
Σ y + ψ x ) dxdy = 0

∫ (ψ dx + φ dy ) − ∫ (ψ dx + φ dy ) = ∫∫ (φ
∂A Θ
Σ x −ψ y ) dxdy = 0

23
Hence

∫ (ψ dy − φ dx ) = ∫ (ψ dy − φ dx ) ,∫ (ψ dx + φ dy ) = ∫ (ψ dx + φ dy )
∂A Θ ∂A Θ

With the above relations we may rewrite (34) as

⎧ (ψ dy − φ dx ) = (ψ dy − φ dx ) − (φ dx −ψ dy )
⎪ ∂∫C Θ∫ ∫Λ
⎨ (35)
⎪ ∫ (φ dy +ψ dx ) = ∫ (ψ dx + φ dy ) + ∫ (φ dy +ψ dx )
⎩∂C Θ Λ

As for (33) in Appendix A, we may similarly show that

⎧ Γi dxi
⎪ ∫ φt dx = − ∫ ψ t dy = − 2 dt
⎪ r ( i ) =ε r ( i ) =ε

⎪ Γ dy
⎪ (i )∫ φt dy = ∫ ψ t dx = − i i
2 dt
⎩ r =ε r ( i ) =ε

Inserting these expressions into (35), we obtain

⎧ dxi dxi d
⎪ ∑Γi dt + ∑Γi dt = dt ∫ (ψ dy − φ dx )
⎪ i ,in i , ∂C
⎨ (36)
dy
⎪− Γ i − Γ i = dy d
⎪⎩ ∑ ∑i (φ dy +ψ dx )
dt i , dt dt ∂∫C
i
i ,in

where ∑ i ,in
is for vortices inside A and ∑ i ,
is over vortices inside  .

The force formula (12) is split here as

dxi dx
L = ρ ∑Γi + ρ ∑ Γi i
i ,in dt i , ou dt
dy dy
D = − ρ ∑ Γi i − ρ ∑ Γi i
i ,in dt i , ou dt

which, when using (36) to replace the first terms on the right hand side, yields

⎧ d dxi
⎪ L = ρ dt ∫ (ψ dy − φ dx ) − ρ ∑ Γi dt
⎪ ∂C i,D
⎨ (37)
⎪ D = ρ d (φ dy + ψ dx ) + ρ Γ dyi
⎪⎩ dt ∂∫C

i,D
i
dt

Here D denotes the region outside of the contour C .

Now, we introduce a downstream boundary x = xd and assume that this is the contour ∂C .
Then with (37) we may write

24
d dy
D=ρ ∫ φ dy + ρ ∑ Γi i
dt x = xd i , xi > xd dt

Through defining ω ( x, y ) = ∑Γ jδ ( x − x j , y − y j ) (where δ is the Dirac function ), we


j
may write

dyi d

i , xi > xd
Γi
dt
= ρ ∫ yω dxdy
dt x > xc

With the identity yω = ∇ ⋅ ( yv, − yu ) + u , and the divergence theorem so that

∫ ∇ ⋅ ( yv, − yu ) dxdy = − ∫ yvdy , we further have


x > xd x = xc

dyi d d

i , xi > xd
Γi
dt
= ρ ∫ ∇ ⋅ ( yv, − yu ) dxdy + ρ ∫ udxdy
dt x > xd dt x > xd
d d
= −ρ
dt ∫x = xd
yudy + ρ ∫ udxdy
dt x > xd
d d
= ρ ∫ φ dy + ρ ∫ udxdy
dt x = xd dt x > xd

where we have used ∫x = xd


yvdy = ∫
x = xd
yφ y dy = − ∫
x = xd
φ dy . Hence

d d
D = 2ρ ∫
dt x = xd
φ dy + ρ ∫ udxdy
dt x > xd

Using the integral form of the y momentum equation for x > xd , we have

udxdy = ∫ ( ρ u 2 + p − ρ u∞2 − p∞ ) dy
d
ρ
dt ∫x > xd x = xc

1 1
and with the Bernoulli equation p = − (u 2 + v 2 ) − φt + p∞ + u∞2 , we have
2 2

φ dy + ρ ∫ (u 2 − v 2 − u∞2 ) dy
d 1
D=ρ
dt ∫x = xd 2 x = xd

Since the above analysis is frame independent, we may choose a frame attached to the vortex street
and therefore φt = 0 on the line x = xd , which is assumed to intersect the vortex street in its

ρ ∫ (u 2 − v 2 − u∞2 ) dy . Now we decompose (u , v) as


1
uniform region. Hence D=
2 x = xd
u = u∞ + uk , v = v∞ + vk = vk , where (uk , vk ) is the induced velocity by the vortex street, then

D=
1
2 ∫x = xd
ρ (uk2 − vk2 ) dy + Dr

25
where Dr = ρ u∞ ∫ u dy . It is obvious that Dr = 0 since the contribution to uk by any
x = xd k

vortex is antisymmetric about the y position of this vortex. Thus we proved (31) in section 3.2.

26

You might also like