You are on page 1of 6

S. M.

RAMODANOV
Department of Theoretical Mechanics
Moscow State University, Vorob’ievy Gory
119899, Moscow, Russia

MOTION OF A CIRCULAR CYLINDER


AND A VORTEX IN AN IDEAL FLUID
Received November 3, 2000 DOI: 10.1070/RD2001v006n01ABEH000163

The motion of a circular cylinder and a point vortex in an unbounded ideal fluid is treated here on the basis of a potential
framework. The formulas for the hydrodynamic force and moment acting upon a cylinder of arbitrary cross section are
obtained. The equations governing the motion of a circular cylinder are derived and partially investigated.

1. Setting up of the problem


Consider a cylinder moving in an unbounded space filled with an ideal fluid. The fluid velocity decays
at infinity and is parallel to a plane orthogonal to an element of the cylinder. Besides the body (the
cylinder) there is a rectilinear filament of strength (−Γ) aligned with an element of the cylinder.
The location of the body in space is well defined by the coordinates x, y of a point O fixed in the
body and by the angle α counted in the counterclockwise direction (Fig. 1). We fix an orthonormal
coordinate frame Oξη to the body. Denote by u, v the components of the velocity of the point O with
respect to the body-fixed frame. Let the angular velocity of the body be denoted by ω.
The motion of the body is governed by the equations

ẋ = u sin α − v cos α, ẏ = u cos α + v sin α,


(1)
α̇ = −ω,

where (ẋ, ẏ) are the components of the velocity of the point O with respect to the inertial frame.
Special cases of motion of cylinders and vortices as well as some references on the subject can be
found in [2].

2. Velocity Field
Denote by C the boundary of a cylinder cross section. The potential that is due to a single vortex of
strength (−Γ) with coordinates (x0 , y0 ) is ϕ∗ = − Γ arg(z − l0 ), l0 = x0 + iy0 . According to [1], [2],

the potential ϕ can be decomposed as follows:

ϕ = uϕ1 + vϕ2 + ωϕ3 + ϕ∗ + ϕ̃.

Here ϕ1 , ϕ2 are the potentials that correspond to translations along the axes Oξ, Oη, respectively.
The potential ϕ3 describes the velocity field that arises due to the rotation of the body with a unit
Mathematics Subject Classification 76C05

REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000 33


S. M. RAMODANOV

angular velocity about the axis perpendicular to the plane of drawing. The impermeable boundary
condition on the contour C and a proper decay at infinity are

∂ϕ4 
= 0, ϕ4 = ϕ∗ + ϕ̃,
∂n C

∂ ϕ̃ ∂ ϕ̃
lim = lim = 0, r 2 = x2 + y 2 .
r→∞ ∂x r→∞ ∂y
Here n is an outward normal vector to C. The explicit expression for ϕ4 is well known if C is a circle.
Anyway, the derivation of this expression using the conformal mapping method is worth mentioning.
Consider a circle C̃ of unit radius on a complex plane of variable ζ. The center of the circle is at the
origin. The motion of the fluid induced by the potential w(ζ) = ϕ + iψ = Γ ln ζ is a pure circulation
2πi
around C̃. This motion could be obtained as well by placing vortices of strengths Γ and −Γ at the
origin and infinity, respectively. Let the inverse image of C̃ under the mapping

z−l
ζ=
1 − ¯lz

be our circle C on a complex plane of variable z. Let W (z) = Φ + iΨ = Γ ln ζ. Note that Ψ = const
2πi
on C, hence for the motion induced by the potential Φ the impermeable boundary condition holds.
The sought potential ϕ4 can be found from the relation

Γ1
w4 (z) = ϕ4 + iψ4 = W + ln z,
2πi

here l is defined through 1/¯l = l0 . The constant Γ1 is a measure of fluid circulation around C.
Theoretically, this constant can take any value but we set Γ1 = −Γ. The latter is justified by the
following: away from the body the velocity field pattern must look approximately the same as that
induced by a single vortex in the absence of the body (i.e. the body is a small disturbance for the
velocity field induced by the vortex). Note that l0 is a single branch point of w4 (z).
Let the complex numbers l, 1/¯l1 locate the center of C and the vortex −Γ, respectively (Fig. 1).
The potential ϕ4 reads
   
ϕ4 = Γ 1
arg(z − l1 − l) − arg z − − l − arg(z − l) (2)
2π ¯l1

Figures 2, 3 illustrate the streamline patterns for Γ = 1 and 10, respectively (the velocity vector of
the center of C is directed towards the vortex).

3. Determination of Forces and Moments


For an inviscid potential flow, the force and moment acting upon a body are known to be given by
Sedov’s formulas [3]  
R = i
d z dϕ = X + iY, M =
d r 2 d ϕ. (3)
dt dt 2
C C
where X, Y are the force components in the inertial frame;
is the density of the fluid. The moment is
referred to the origin of the inertial frame. A detailed derivation of (3) can be found in, for example, [1].
For the problem under consideration, the formulas (3) need be corrected.
Let us use, in accordance with [1], the principle of momentum. Consider concentric circles C1
and C2 the centers of which coincide with the vortex (Fig. 4). Let the radii r1 and r2 of the circles be

34 REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000


MOTION OF A CIRCULAR CYLINDER AND A VORTEX IN AN IDEAL FLUID

so chosen that they do not intersect C. We make a cut AB such that in the area D enclosed by C, C1 ,
C2 , AB the potential is a single-valued function. The impulse of the fluid within D can be written as
 
∂ϕ ∂ϕ
P = Px + iPy =
dτ + i
dτ.
∂x ∂y
D D

Here Px , Py are the projections of the impulse in the inertial frame. Using the Green-Gauss formula
and integrating by parts, we get
  
P = Px + iPy =
ϕ dy − i
ϕ dx = i
z dϕ.
∂D ∂D ∂D

The values of ϕ on the two sides of AB differ by Γ and thus the corresponding integrals vanish. We
have   
P = Px + iPy = i
z dϕ − i
z dϕ − i
z dϕ.
C2 C1 C

Denote by RC , RC1 , RC2 the hydrodynamic forces applied to the circles C, C1 , C2 , respectively. The
time rate of change of the impulse is known to be
  
dP = i
d z dϕ − i
d z dϕ − i
d z dϕ = −RC − RC1 + RC2 . (4)
dt dt dt dt
C2 C1 C

Let us prove that the difference RC2 − i
d z dϕ vanishes as r2 → ∞. The pressure can be found
dt
C2
through the use of the Cauchy–Lagrange formula

∂ϕ
v 2
p = −
− + F (t).
∂t 2
Therefore   
RC2 = i p dz = −i

∂ϕ
dz − i
v 2 dz.
∂t 2
C2 C2 C2

Then     
d d z dϕ + dϕ ∂ϕ  2 
z dϕ = zd = (vx + ivy ) dϕ − + vx + vy2 dz.
dt dt dt ∂t
C2 C2 C2 C2 C2

The result is
   
     2
vx2 − vy2 vy − vx2
RC2 − i
d z dϕ =
 vx vy dx − dy  + i
− vx vy dy + dx . (5)
dt 2 2
C2 C2 C2 C2 C2
 
Since v = O r12 as r2 → ∞, we see that both integrals in the right-hand side of (5) vanish as
r2 → ∞. 
Find the limit of RC1 − i
d z dϕ as r1 → 0. Let the origin of the inertial frame coincide with
dt
C1
the center of C1 (Fig. 4). For a fixed time t we have
 
Γ y −Γ x
vx + ivy = + ṽx + i + ṽy .
2π x2 + y 2 2π x2 + y 2

REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000 35


S. M. RAMODANOV

Here ṽx , ṽy are analytic functions of x, y in a vicinity of the origin. Let ṽx (0, 0) = u0 , ṽy (0, 0) = v0 .
Using (5), we get
 
  
− Γ

RC1 − i
d z dϕ = (−v0 y + u0 x) dx − (−u0 y − v0 x) dy  −
dt 2πr12
C1 C2 C2
 
 


− (v0 x + u0 y) dx − (−v0 y + u0 x) dy  + . . .,
2πr12
C2 C2

where unwritten terms vanish as (x, y) → 0. Consequently, the sought limit value is −Γ
v0 + iΓ
u0 .
Letting r2 → +∞ and r1 → 0 in (4), we finally obtain the expression for the force acting on C

RC = i
d z dϕ + Γ
v0 − iΓ
u0 . (7)
dt
C

Comparing this formula with (5), we see that the presence of a vortex results in additional terms in
the formula for the force.
Now let us calculate the moment about the origin of the inertial frame acting upon C. The
angular momentum of the fluid in D reads
     
K=

∂(xϕ) (∂yϕ)
− dτ = −
ϕ(x dx + y dy) =
r 2 dϕ −
r 2 dϕ −
r 2 dϕ. (8)
∂y ∂x 2 2 2
D ∂D C2 C1 C

The forces acting upon C1 and C2 produce no moment since their lines of action pass through the
origin. Hence the moment applied to C is M = −dK/dt. Consider the time rate of change of the
integrals in (8). The equality 
2
lim d
r dϕ = 0
r2 →+∞ dt 2
C2

holds because the flow velocity decays at infinity approximately as r12 . The time rate of the integral
along C1 is
     
d
r 2 dϕ =
d r 2 dϕ =
xv 2 + yv v  dx +
yv 2 + xv v  dy.
x y x y y x
dt 2 dt 2
C1 C1 C1 C1

It can be easily shown, by use of (6), that the expression in the right-hand side becomes zero as
r1 → 0. Differentiating (8) and letting r2 → +∞, r1 → 0, we obtain the expression for the moment
acting on C 
M =
d r 2 d ϕ. (9)
dt 2
C

This equation is identical in form to (3) but note that now the moment is calculated about the point
that coincides with the vortex.
Calculate the integral in the expression for the force (7). Again consider a complex plane of
variable ζ and let C̃ be a circle of unit radius with the center at the origin. A conformal mapping of
the exterior of C̃ on the exterior of C is

z = f (ζ) = kζ + k0 + k1 ζ −1 + . . .

36 REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000


MOTION OF A CIRCULAR CYLINDER AND A VORTEX IN AN IDEAL FLUID



here k = dz  > 0. By use of this relation, (7) can be rearranged as follows:
dζ ∞

z dϕ4 =
C   
  Γ arg(ζ − l ) − arg ζ − 1/¯l  − arg ζ  =
= kζ + k0 + k1 ζ −1 + . . . d 1 1

C̃ 
  
 kζ + k0 + k1 ζ −1 + . . . kζ + k0 + k1 ζ −1 + . . . kζ + k0 + k1 ζ −1 + . . . 
= Γ  dζ − dζ − dζ  =
2πi ζ − l1 ζ − 1/¯l1 ζ
 C̃ C̃  C̃
k
= Γ 2πi(f (l1 ) − 1 ) + 2πik1 ¯l1 − 2πik0 = Γ (f (l1 ) − k0 ).
2πi l1
(10)
Note that replacing ϕ by ϕ4 in (7), we find that part of the total force applied to C which is induced
by the vortex and circulation. Of course, an additional influence of the ambient flow on the body
appears through the added-mass effect, which is well known and requires no special considerations.
Finally, in view of (7) and (10), the force induced by the circulation and the vortex is seen to be

R = −iΓ
d k0 + iΓ
d f (l1 ) + Γ
v0 − iΓ
u0 , (11)
dt dt
The first term in the right-hand side is due to the circulation. This force is defined by the velocity of
the point k0 fixed in the body. It is also seen from (11) that the body is subjected to a force (the last
two terms) that depend on the velocity of the vortex and of the point f (l1 ).

4. Equations of Motion of a Circular Cylinder


Again let C be a circle of unit radius. Setting k = 1, k0 = l, kj = 0 (j ∈ N) in (11), we see that the
force applied to the body
R = −iΓ
d l + iΓ
d l1 + Γ
v0 − iΓ
u0 , (12)
dt dt
depends on the velocities of the points l1 , 1/¯l1 and l (Fig. 1).
Using (9), one can check that the moment applied to C is zero. This immediately follows from
the fact that the fluid is ideal and thus there is no friction.
We assume now that the axes of the frame Oξη are aligned with the axes of the inertial frame.
Let 1/¯l1 = ξ2 + iη2 be the coordinates of the vortex with respect to Oξη. Hence the coordinates of
the point l1 = ξ1 + iη1 are
ξ2 η2
ξ1 = 2 2
, η1 = 2 . (13)
ξ2 + η2 ξ2 + η22
The velocity of the point l1 relative to the point O has the components ξ˙1 , η̇1 . They could be found
from (13) and the relations 
∂ϕ(ξ, η) 
ξ̇2 = −u +  ,
∂ξ 
ξ=ξ2 ,η=η2

∂ϕ(ξ, η) 
η̇2 = −v +  .
∂η 
ξ=ξ2 ,η=η2
Here    
ξ η η − η1 η
ϕ(ξ, η) = −u 2 − v 2 + Γ arctg − Γ arctg .
r r 2π ξ − ξ1 2π ξ

REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000 37


S. M. RAMODANOV

After rearrangement, the equations of motion of the body, in view of (5), take the form

u̇ = µv − µη̇1 + µη̇2 , v̇ = −µu + µξ̇1 − µξ˙2 ,


r 4 η1
ξ̇1 = −u(r 4 + η12 − ξ12 ) + 2vη1 ξ1 + Γ 2 ,
2π r − 1 (14)
r 4 ξ1
η̇1 = −v(r 4 − η12 + ξ12 ) + 2uη1 ξ1 − Γ 2 .
2π r − 1

Here r 2 = ξ12 + η12 , µ = Γ


/a. The constant coefficient a involves the added mass of the body. In view
of (13), the first two equations of (14) can be written as (here, and in the sequel, the index 1 in the
notation ξ1 , η1 is dropped)

u̇ = µvr 4 + λr 2 ξ, v̇ = −µur 4 + λr 2 η. (15)

Here λ = Γµ/2π. Equations (14), (15) have a first integral


2 +v 2 )/λ
(1 − r 2 )e(u = const.

The time derivative of uξ + vη is a function of r 2 , u2 + v 2 , uξ + vη. Hence, using (15) and the first
integral, it is feasible to obtain a rather cumbersome second-order differential equation in r 2 . For any
0  c < 1 the initial velocity of the point O can be so chosen that during the motion r 2 is equal to c.
In this case, the point O and the vortex proceed along concentric circles. For other initial data, as it
may be seen from Fig. 5, 6, the path of the vortex (the solid line) and the path of the point O (the
dashed line) may look rather fancifully.

References

[1] N. E. Cochin, I. L. Kibel, N. V. Roze. Theoretical Hy- [3] L. I. Sedov. Two-Dimensional Problems in Hydro- and
drodynamics. . 1953. P. 394. (in Russian). Aeromechanics. Ed. 2-th. . Gostechizdat. 1950. (in
[2] H. Lamb. Hydrodynamics. Ed. 6-th. New York: Dover Russian).
Publications. Inc. 1945.

38 REGULAR AND CHAOTIC DYNAMICS, V. 6, No. 1, 2000

You might also like