Added Mass of A Circular Cylinder Oscillating in A Free Stream

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Downloaded from http://rspa.royalsocietypublishing.

org/ on January 13, 2016

Added mass of a circular


cylinder oscillating
in a free stream
rspa.royalsocietypublishing.org Efstathios Konstantinidis
Department of Mechanical Engineering, University of Western
Macedonia, Bakola and Sialvera, Kozani 50100, Greece

Research The fundamental understanding of the added mass


phenomenon associated with the motion of a solid
Cite this article: Konstantinidis E. 2013
body relative to a fluid is revisited. This paper
Added mass of a circular cylinder oscillating in focuses on the two-dimensional flow around a circular
a free stream. Proc R Soc A 469: 20130135. cylinder oscillating transversely in a free stream. A
http://dx.doi.org/10.1098/rspa.2013.0135 virtual experiment reveals that the classical approach
to this problem leads to a paradox. The inertial force
is derived afresh based on analysis of the motion
Received: 25 February 2013
in a frame of reference attached to the cylinder
Accepted: 9 May 2013 centroid, which overcomes the paradox in the classical
formulation. It is shown that the inertial force depends
not only on the acceleration of the cylinder per
Subject Areas:
se, but also on the relative motion between body
fluid mechanics and fluid embodied in a parameter called alpha, α,
which represents the ratio of the maximum transverse
Keywords: velocity of the cylinder to the free-stream velocity;
fluid mechanics, fluid–structure interaction, the induced inertial force is directionally varying
potential flow and non-harmonic in time depended on the alpha
parameter. It is further shown that the component
of the inertial force in the transverse direction is
Author for correspondence: negligible for α < 0.1, increases quadratically for α <
Efstathios Konstantinidis 0.5, and tends asymptotically to the classical result as
e-mail: ekonstantinidis@uowm.gr α → ∞, i.e. in still fluid.

1. Introduction
Whenever acceleration is imposed on a fluid by
acceleration of a body moving through the fluid, or
acceleration of the fluid relative to the body, additional
inertial forces act on the body due to the ‘added mass’
effect. The calculation of these inertial forces has been of
fundamental interest in fluid mechanics for more than
a century. It can also be of considerable significance
in some engineering applications involving lightweight
structures submerged in dense fluids. One such example
is the ‘vortex-induced vibration’ of offshore structures
encountered in exploration and extraction of petroleum
from sea beds; marine cables, pipelines, risers, platforms,

2013 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

etc., can be excited into significant vibration by sea currents and waves and dealing with its
2
consequences has provided impetus and funding for concentrated research on the problem over
the last few decades. Some relatively recent fundamental studies have highlighted the importance

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
of added mass in vortex-induced vibration and increased interest in this topic [1–3].
The scope of this work is to revisit the fundamental understanding of the added mass
effect with particular attention to the ‘deceivingly’ simple case of a circular cylinder oscillating
transversely in a free stream. A review of the literature in §2 reveals that some fundamental
aspects of the added mass notion may have been ignored in the classical formulation of this
problem. The inconsistency of the classical formulation with the physical world is demonstrated
through a virtual experiment in §3. A new formulation of the added mass in a frame of reference
moving with the cylinder is given in §4 to derive the inertial force acting on a cylinder oscillating
transversely to a free stream from potential flow theory. However, a purely mathematical treatise
of the flow potential is beyond the objectives of the present work. The paper concludes with some
remarks from the present work in §5.

2. Literature review
The notion of added mass, also referred to as hydrodynamic mass or virtual mass in the literature, can
be found in textbooks of classical hydrodynamics [4,5], but according to some sources, it was first
introduced by Dubua back in 1776 [6, ch. 1]. A survey of the literature shows that added mass is
customarily—but not uniquely—derived from the kinetic energy associated with the motion of
the fluid. The formulation below is adapted from Brennen [7].
The kinetic energy of the fluid, EK , can be expressed in tensor notation as

1
EK = ρ ui ui dV, (2.1)
2 V
where V is the entire volume of fluid, ui (i = 1, 2, 3) represent the spatial components of the fluid
velocity in this volume, and repeated indices imply a summation. It is assumed throughout that
the fluid is incompressible with density ρ. Any increase in the kinetic energy of the fluid implies
that additional work must be done on the fluid to supply the extra energy, and vice versa.
Consider the motion of a non-deformable body through an unbounded fluid otherwise at rest
is a rectilinear translation at velocity, U. The kinetic energy can be conveniently expressed as

1 ui u i
EK = ρIU2 , where I = dV. (2.2)
2 V U U
If the velocity of the body is steady, the kinetic energy of the fluid remains constant in time even
for unsteady and/or turbulent flows (in a statistical sense). If the body is accelerated the velocity
at each point in fluid relative to the body varies in direct proportion to the change in U. The rate
of additional work required is equal to the rate of change of the kinetic energy with time. This
comes to the effect that an additional inertial force, FI , is experienced by the body such that the
work done by the body −FI U is equal to dEK /dt, i.e.
dU
FI = −ρI . (2.3)
dt
The inertial force has the same form and sign as that required to accelerate the mass, m, of the
body itself, namely m (dU/dt). Consequently, it is rational to consider the mass of fluid given by
ρI as an added mass, denoted mA , of fluid being accelerated ‘by’ the body—but not ‘with’ the body.
This subtle distinction has some important ramifications which are addressed further below. As
pointed out by Brennen [7, p. 3], ‘there is no such identifiable fluid mass; rather all of the fluid is
accelerating to some degree such that the total kinetic energy of the fluid is increasing’.
The above statement reveals another subtlety in the analytical description of incompressible
flows: in theory, a change in pressure on the body surface must be instantly radiated throughout
the entire fluid domain. This can be problematic when it comes to the computation of volume
integrals over the entire fluid domain such as that defining the kinetic energy in (2.1). In this
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

respect, it may worth drawing the attention to the indeterminacy of the integral defining Darwin’s
3
drift mass which, coincidentally, is always equal to the added mass for incompressible flows [8].
The point is made to illustrate that the definition of the added mass needs be modified if the fluid

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
far from the body is not at rest as assumed hitherto.
Now consider the motion of a body relative to an unbounded fluid stream (‘free stream’)
of steady velocity, U∞ , far from the body. In this case, the kinetic energy within an infinitely
large domain becomes infinite and the above formulation of the added mass is meaningless. The
appropriate fluid velocities, ui , to be employed in defining the kinetic energy in (2.1) are those
relative to the free-steam velocity,

1
EK = ρ (u − U∞ )i (u − U∞ )i dV. (2.4)
2 V
Then, the integral defining EK becomes determinable. It can be shown that, following the same
procedure as above, this Galilean transformation leads to no alteration in the inertial force since
d dU
FI = −ρI (U − U∞ ) = −mA , (2.5)
dt dt
which is the same result as (2.3). It is assumed above that U and U∞ are aligned but even if the two
velocities are not collinear, it can readily be verified that the same result is obtained by employing
vectors. This is a deceivingly simple result which is very often employed as the starting point
to define added mass in the literature, rather than the opposite. In fact, this has been the root of
much confusion regarding the added mass as its meaning is often misinterpreted as

inertia force = added mass × body acceleration

irrespectively of the fluid motion relatively to the body. The above expression is incorrect unless
the fluid far from the body is at rest. For all other cases, it is not in accord with the physical
world as is demonstrated through a virtual experiment in §3. It is worth noting here the words of
Sarpkaya [2]: ‘the added mass is not a concentrated mass attached to the centroid of the body. It is
distributed throughout the fluid set in motion by the body. Thus, its magnitude and centroid
change with time as the intensity and distribution of the kinetic energy of the fluid change
with time.’
For completeness, it is useful to generalize the definition of the added mass. For a body of
arbitrary shape moving within a fluid, there are in general three translational accelerations, U̇j (j =
1, 2, 3) and three rotational accelerations U̇j (j = 4, 5, 6), where the overdot denotes differentiation
with respect to time. The added mass expresses in a unified way the linear forces and moments,
Fi (i = 1, 2, . . . , 6), imposed on the body due to all six possible components of acceleration,

Fi = mij U̇j . (2.6)

For the general case of any body in arbitrary motion through a fluid, there are 36 different added-
mass components, mij (i, j = 1, 2 . . . 6). Under specific conditions, some of these components can
be related, e.g. due to simplifications introduced by geometric symmetries. For a body with three
planes of symmetry, the number of independent added mass components reduces to 12. Further
simplifications arise for potential flow in which case the added mass matrix is symmetric ([9]
cited in [7]). The general expression in (2.6) is the appropriate starting point for the study of
the dynamic behaviour of a body moving within a fluid, rather than (2.3) as there is a risk of
overlooking terms that may be significant [10].
The above formulation to derive the added mass from the kinetic energy of the fluid makes
no assumption about the nature of the flow (rotational or irrotational, laminar or turbulent, etc.).
It follows from the foregoing that all characteristic flow structures and processes generated by
the motion of a body within a viscous fluid, such as boundary layers, free shear layers, wakes,
formation and shedding of concentrated vorticity in the wake, contribute to the kinetic energy
of the fluid and their evolution in time is what determines the inertial force due to the added
mass. Hence, knowledge of the entire flow field over time is required to evaluate the kinetic
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

energy when a body moves within a real fluid; this is one of the most important and challenging
4
problems in fluid mechanics [11]. Particularly in the case of unsteady separated turbulent flows,
the problem is even more complicated.

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
Evaluation of the inertial forces through the general expression in (2.6) require as a
precondition that the flow induced by each component of the body motion or velocity in a fluid
are linearly superposable. This is generally not true because the equations of fluid motion (e.g.
Navier–Stokes) are nonlinear. However, there are two models of fluid flow which do satisfy
the superposability condition, namely the potential flow for infinitely large Reynolds numbers
and the Stokes flow for asymptotically small Reynolds numbers [7]. Quoting from the latter
source (p. 9): ‘Only in these limiting cases can the added mass matrix be regarded as an exact
representation of the relation between fluid inertial force and body acceleration. In other types of
flow it could however be regarded as a reasonable approximation.’
In the framework of inviscid potential flow, analytical expressions have been developed to
evaluate added mass components for specific body geometries in different kinds of motion, which
can be found in the literature [7,10]. The problem becomes much more difficult in viscous flows
because of the effect of vorticity generated on the body surface and preexisting within the fluid.
However, a number of studies have demonstrated that added-mass effects are independent of the
nature of the flow. First and foremost, Basset [12] showed that the general equations of impulsive
motion of a viscous fluid are the same as those of a perfect fluid based on a very simple but strict
reasoning. Chang [13] showed in a rigorous manner that the effect due to the body’s acceleration
can be separated from the contribution from vorticity on the force acting on the body. Chang [13]
also showed that the potential flow has a geometric influence on the contribution of vorticity
attached on the body surface and within the flow to the drag and lift forces, i.e. viscous forces are
not free from inertial effects. Leonard & Roshko [1] argued that added mass definitions obtained
for irrotational flow are more generally applicable to incompressible flow because an incremental
change in velocity instantaneously generates a potential velocity field proportional to that velocity
change which is superimposed on the existing velocity field. Mougin & Magnaudet [14], based
on the mathematical treatment of the equations of motion by Howe [15], showed that the linear
and angular momentum balances for a body moving within a viscous rotational flow at rest at
infinity can be cast in a form which comprises exactly the added-mass components as for the
corresponding motion within an inviscid fluid; the contributions from vorticity appear as separate
terms in the equations. Mougin & Magnaudet [14] further argued that added-mass effects are
independent of the body surface, i.e. interchanging slip and no-slip boundary conditions makes
no difference to the inertial forces. Wakaba & Balachandar [16] firmly established through direct
numerical simulations the independence of the added mass coefficient to the acceleration number
and to the state of flow prior to acceleration for flows around a rigid sphere in relative motion
to a viscous fluid; the explanation provided by the latter authors is that the potential flow part
establishes instantaneously as in the case of an ideal fluid (no memory effect).
A critical review of previous studies discussed in the preceding paragraphs reveals three
important aspects: (i) added mass can be properly defined in both ideal inviscid and real viscous
flows but exact evaluations can be made only in the former case, (ii) inviscid flow added-mass
effects are inherent in viscous flows for the same kind of body motion relative to the fluid
independently of the nature of the flow (i.e. for instantaneous Reynolds numbers from zero to
infinity), and (iii) the correct way to evaluate added masses is in a frame of reference fixed with
the body moving within a fluid otherwise at rest far from the body. The last point appears not to
be understood clearly. The required forces in a fixed frame of reference have to be obtained by
‘projecting’ the absolute velocity field from the ‘moving’ to the fixed frame of reference [13,14].

3. The added mass paradox: a virtual experiment


This work is concerned with two-dimensional flow around a circular cylinder oscillating
transversely in a free stream. This is one of the conceptually most simple configurations to study
the added mass effect, but not as simple as it may originally seem. Surprisingly, no solutions
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

(a) (b)
5

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
U U

Figure 1. Streamlines of potential flow around a circular cylinder accelerating through a free stream of steady velocity for (a)
rectilinear and (b) orbital motion.

specific to this flow configuration can be found in textbooks or other probable sources in the
literature. The derivation of the inertial force is postponed to §4.
The following example is adapted to the case of a circular cylinder from one with a prolate
spheroid given by Imlay [10]. Consider a circular cylinder in rectilinear motion through an ideal
inviscid fluid in the direction of the x-axis, and experiencing acceleration, U̇. The potential flow
would have streamlines of the sort shown in figure 1a. Consider now the cylinder is at the same
position having the same acceleration as before but due to orbital motion at a fixed angle and
rate of drift. The nature of a real flow is to align with the instantaneous direction of the orbiting
cylinder; this can be clearly seen in flow visualization photographs of the wake behind an orbiting
cylinder (see [17, fig. 3]). The streamlines for the corresponding potential flow would be those
depicted in figure 1b. Although the actual mass of the body would react in the same way in
both cases, the added mass effect, i.e. the inertial force in the x-direction, would be different
because of the difference in flow patterns attending the two different motions. Recalling that a
given flow pattern requires a unique velocity potential to produce it; that the kinetic energy is
derivable from the velocity potential; and that the added mass effects are the result of changes
in the kinetic energy of the fluid; it reasonable to expect different added mass effects in the two
different flow patterns [10]. The above example illustrates that the added mass does not depend
on the acceleration of the body per se, so the added mass effect may differ under two sets of
circumstances where the acceleration is the same. It is clear that, the change in the magnitude of
the linear velocity would cause an inertial force in the x-direction in the first case, whereas the
same acceleration in the second case would not produce any inertial force since the magnitude of
the linear velocity of the cylinder and the total kinetic energy of the fluid remain constant in time.
Now, consider the flow around a circular cylinder oscillating transverse to a free stream.
It follows from the foregoing that it is very unlikely that the added mass effect is the same
as if the cylinder is oscillating in still fluid. However, that is invariably what is assumed in
the classical approach to this problem in the literature. Let the cylinder undergo sinusoidal
oscillations transverse to the free stream, its motion being described as a function of time by
Y(t) = A sin ωt, where A is the amplitude of oscillation and ω is the cyclic frequency. According
to the classical formulation, the potential-flow added mass depends only on the shape of body
and the direction of motion [18, p. 567]. Hence, the only component of the inertial force due to the
added-mass effect acts in the transverse y-direction,

d2 Y
FIy = −mA = mA ω2 A sin ωt, (3.1)
dt2
where mA is the added mass. For two-dimensional potential flow around a body with geometry
having three planes of symmetry, the only independent components in the added mass matrix
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

are the diagonal terms. For the circular cylinder, in particular, the two components (in Cartesian
6
coordinates) associated with translational acceleration are identical and equal to the fluid mass
displaced by the cylinder. Suppose one carries out experiments where A and ω can be varied at

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
will, and the force acting on the cylinder is measured. A set of parameters is employed for which
the amplitude is kept negligibly small but finite, and the frequency is varied from zero to a very
large value. It is obvious that the flow pattern around the cylinder in all experiments would be
similar to that around a stationary cylinder since the amplitude is negligibly small; hence, the
kinetic energy of the fluid would also remain unaffected, and there would be no induced inertial
force acting on the cylinder. However, the measured force is expected to vary in direct proportion
to ω2 according to the classical formulation in (3.1). This is obviously impossible and is called here
the added-mass paradox.

4. Derivation of the added-mass force on an oscillating cylinder


Consider a circular cylinder oscillating transversely in a free stream of steady uniform velocity,
U∞ . The motion of the cylinder induces a time-dependent relative velocity U(t) = U∞ i − Ẏ(t)j
where i and j are the orthonormal basis in a fixed frame of reference in Cartesian coordinates {x, y},
Y(t) describes the cylinder’s motion and the overdot denotes a derivative with respect to time. The
relative velocity vector can also be expressed in a frame of reference moving with the cylinder as
U(t) = U(t)eη where eη is the basis vector aligned with the instantaneous direction of the cylinder’s
velocity relative to the fluid in polar coordinates {η, θ} as shown in figure 2. In essence, the cylinder
motion vis à vis the free stream can be decomposed into an equivalent translation plus a rotation
of the cylinder about its centre. It should be noted that the cylinder does not literally rotate in
the fixed frame of reference. The acceleration can be obtained by differentiation of the relative
velocity with respect to time in the frame of reference attached to the cylinder,
dU
U̇(t) = = U̇ eη + Uθ̇ eθ , (4.1)
dt
where θ is the angle defined in figure 2, θ̇ is the angular acceleration of the frame of reference, and
eθ the corresponding basis vector.
Each of the two terms on the right-hand-side of (4.1) is associated with independent
components of the added mass matrix due to translational and rotational accelerations and
corresponding inertial forces acting on the body in the general case of bodies with arbitrary
geometry in two dimensions. For a circular cylinder, the potential-flow added mass for translation
equals the fluid mass displaced by the cylinder, whereas the added mass for rotation about its
centre is zero. Hence, only the first term contributes to the inertial force induced by the added
mass effect, i.e.
F I = mA U̇ = mA U̇eη , (4.2)

where mA is the added mass of a circular cylinder associated with translation (the only non-
zero term). Equation (4.2) shows that the induced inertial force is directionally varying (aligned
with the instantaneous relative flow direction) and its magnitude varies in time according to the
derivative of the relative velocity magnitude, whatever that may be. Hence, the result in (4.2)
while specific to the circle is generally applicable to all kinds of oscillation, e.g. sinusoidal or non-
sinusoidal, with a single frequency or multiple components, and even more broadly to all kinds
of accelerations, e.g. impulsive or random.
If the cylinder oscillation is assumed for simplicity to be sinusoidal in time, Y(t) = A sin ωt, the
magnitude of the relative velocity vector is given by

U(t) = U∞ 1 + α 2 cos2 ωt, (4.3)

where α = ωA/U∞ is the ratio of maximum transverse velocity of the cylinder to the free-stream
velocity, called the alpha parameter hereafter. The inertial force in the moving frame of reference
can be obtained by differentiation of (4.3) and substitution of the result in (4.2). Projection of the
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

7
U•
q

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
y U

x =U• t

Figure 2. Schematic showing a circular cylinder oscillating transverse to a free stream in a frame of reference moving with the
cylinder.

1.0
0.8
0.6
G(t)
0.4
0.2

0 0.25 0.50 0.75 1.0


t/T

Figure 3. Variation of the G function with dimensionless time over a cycle of oscillation: solid line, α = 0.2; dashed line,
α = 0.5; dotted line, α = 1.0; dashed-dotted line, α = 2.0.

inertia force back into the fixed frame of reference can be accomplished via the transformation
eη = cos θ i + sin θ j. After some mathematics, the following result is obtained for the inertial force
in the y-direction:
d2 Y
FIy (t) = mA 2 G(t; α), (4.4)
dt
where G(t; α) is a dimensionless function of time which depends solely on the alpha parameter.
This result demonstrates that the inertial force due to added mass is not only a function of the
body geometry and the direction of motion, but also of the relative motion, embodied in the
alpha parameter, in contrast to the classical formulation. The new result in (4.4) differs from the
classic one in (3.1) only in the term G(t; α). However, this is not a simple correction. The analytical
form of the G function is
α 2 cos2 (ωt)
G(t; α) = . (4.5)
1 + α 2 cos2 (ωt)
Figure 3 shows the temporal variation of the G function for different values of α. The G function
is positively defined varying between minimum and maximum values of zero and (1 + α −2 )−1 ,
respectively. G(t) is a non-harmonic function of time with a fundamental frequency at twice the
cylinder’s oscillation frequency. Hence, the added mass even for the simple case of potential flow around
a circular cylinder undergoing sinusoidal oscillation transverse to a free stream gives rise to a complex
non-harmonic inertial force in the same direction.
As a side note, the inertial force induced in the transverse direction, FIy (t), comprises
main components at the fundamental and the third harmonic of the cylinder’s oscillation
frequency since it approximately varies according to the following trigonometric relationship
sin(ωt) cos2 (ωt) = 14 [sin (ωt) + sin (3ωt)].
The contribution of the added mass to the transverse force for an oscillating cylinder can be
appreciated from the variation of Gmax = max{G(t)} as a function of the alpha parameter, which is
shown in figure 4. For α < 0.1, the inertial force in the transverse direction is practically negligible
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

1.0
8
0.8

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
0.6

Gmax
0.4

0.2

0
0.01 0.1 1 10 100
a

Figure 4. Variation of Gmax with α.

since Gmax ≈ 0. For 0.1 < α < 10, Gmax increases rapidly tending asymptotically to unity as α → ∞.
For α  0.5, the inertial force increases quadratically with the alpha parameter since Gmax ∝ α 2 .

5. Concluding remarks
In this paper, a new formulation is employed to derive the inertial force on a body oscillating
transversely to a free stream due to the added-mass effect. The formulation is based on analysis
of the body motion in a frame of reference fixed to the centroid of the cylinder moving through
still fluid as this is a prerequisite for the proper definition of the added mass from the kinetic
energy of the fluid. It is shown that the inertial force depends not only on the body geometry and
the direction of its motion, but also on the flow pattern induced by the motion of the body relative
to the free stream. This result stands in sharp contrast to the classical assumption that the inviscid
potential-flow force is simply equal to the product of a shape- and motion-dependent constant
added mass times the acceleration of the body. The latter assumption is frequently employed
in the literature but is not physically consistent as demonstrated by a virtual experiment. This
is so because the added mass resides outside the surface of the body, whereas in the classical
formulation, it is replaced by an equivalent mass attached to the centroid of the body.
For a circular cylinder in sinusoidal oscillation transverse to a free stream, the dependency
of the inertia force on the relative motion between body and fluid can be embodied in a single
parameter called alpha, α = ωA/U∞ . Even in this simple case, the added-mass for potential flow
gives rise to a complex directionally varying non-harmonic in time inertial force. Projection of the
inertial force to the transverse direction yields a component which is negligible for α < 0.1 but
its contribution increases quadratically for α < 0.5. It can be readily verified that the new result
in (4.4) recovers the classical result in still fluid (α → ∞) and dispels the added mass paradox in
the presence of a free stream; i.e. the paradox is not a shortcoming of potential flow per se but of
the limitation of an inviscid fluid to meet the boundary conditions of a viscous fluid.
A key finding of the present study is that a better way to conceptualize the inertia force due to
the added mass is

inertia force = added mass × added-mass acceleration.

For the specific configuration considered here, the added-mass acceleration is equal to the
acceleration of the oscillating body d2 Y/dt2 times the dimensionless function G(t; α), whose
analytical form can be worked out. For geometries other than a circle, the analytical calculation of
the inertial force becomes more complicated because it comprises, in general, the contributions
due to both translational and rotational accelerations and thereby different added mass
components.
Based on the outcome of the present study, the physical meaning of the added mass is the
following: as a body moves through a fluid, it displaces the fluid particles permanently in the
Downloaded from http://rspa.royalsocietypublishing.org/ on January 13, 2016

direction of its motion; if the body accelerates part of the fluid close to the body will also accelerate
9
while being pushed aside by the body. Thence, an additional inertial force is required to accelerate
this part of the fluid. Even though the part of the fluid being accelerated by the body is infinitely

rspa.royalsocietypublishing.org Proc R Soc A 469: 20130135


..................................................
widely spread in theory, this does not matter to the physical world; it has been demonstrated in
a few studies that contributions of flow structures to the forces acting on the body decay rapidly
with distance from the body [13,19]. This development is consistent with the equivalence of the
added mass and the drift mass. The interested reader may wish to consult the resolution of the
indeterminacy of the drift mass integral provided by Yih [8]. By virtue of its definition and its
physical meaning, added mass is a positively defined quantity. The total added mass cannot
be negative even though the drift mass may exhibit a local reflux. However, the component of
the induced inertial force in the direction of body motion can become negligible under certain
conditions as is shown in the present work.
Added mass is a generic property in fluid mechanics and its fundamental understanding has
far-reaching implications in problems involving flow–structure interactions. In particular, added
mass has a significant effect on the vortex-excited oscillation of cylindrical structures when the
density of the structure is comparable with the fluid density, which has been the subject of much
debate in the literature [2]. It is expected that the present study will shed new light on this
outstanding issue.

References
1. Leonard A, Roshko A. 2001 Aspects of vortex-induced vibration. J. Fluids Struct. 15, 415–425.
(doi:10.1006/jfls.2000.0360)
2. Sarpkaya T. 2004 A critical review of the intrinsic nature of vortex-induced vibrations. J. Fluids
Struct. 19, 389–447. (doi:10.1016/j.jfluidstructs.2004.02.005)
3. Leong CM, Wei T. 2008 Two-degree-of-freedom vortex-induced vibration of a pivoted cylinder
below critical mass ratio. Proc. R. Soc. A 464, 2907–2927. (doi:10.1098/rspa.2007.0166)
4. Lamb H. 1932 Hydrodynamics. Cambridge, UK: Cambridge University Press.
5. Milne-Thomson LM. 1938 Theoretical hydrodynamics. London, UK: Macmillan & Co Ltd.
6. Korotkin AI. 2009 Added masses of ship structures. Berlin, Germany: Springer.
7. Brennen CE. 1982 A review of added mass and fluid inertial forces. Tech. Report no. CR 82.010,
Naval Civil Engineering Laboratory, Port Hueneme, CA, USA.
8. Yih C-S. 1997 Evolution of Darwinian drift. J. Fluid Mech. 347, 1–11. (doi:10.1017/S002211
209700654X)
9. Yih C-S. 1969 Fluid mechanics. New York: McGraw-Hill Book Co.
10. Imlay F. 1961 The complete expressions for ‘added mass’ of a rigid body moving in ideal fluid.
Technical Report no. 1528, Hydromechanics Laboratory, David Taylor Model Basin.
11. Sarpkaya T. 2001 On the force decompositions of Lighthill and Morison. J. Fluids Struct. 15,
227–233. (doi:10.1006/jfls.2000.0342)
12. Basset AB. 1888 On the motion of a sphere in a viscous liquid. Proc. R. Soc. Lond. A 179, 43–63.
(doi:10.1098/rsta.1888.0003)
13. Chang C-C. 1992 Potential flow and forces for incompressible viscous flow. Proc. R. Soc. Lond.
A 437, 517–525. (doi:10.1098/rspa.1992.0077)
14. Mougin G, Magnaudet J. 2002 The generalized Kirchhoff equations and their application to
the interaction between a rigid body and an arbitrary time-dependent viscous flow. Int. J.
Multiphase Flow 28, 1837–1851. (doi:10.1016/S0301-9322(02)00078-2)
15. Howe MS. 1995 On the force and moment on a body in an incompressible fluid, with
application to rigid bodies and bubbles at high and low Reynolds numbers. Q. J. Mech. Appl.
Math. 48, 401–426. (doi:10.1093/qjmam/48.3.401)
16. Wakaba L, Balachandar S. 2007 On the added mass force at finite Reynolds and acceleration
numbers. Theor. Comput. Fluid Dyn. 21, 147–153. (doi:10.1007/s00162-007-0042-5)
17. Wu J, Sheridan J. 1997 The wake of an orbiting cylinder. J. Fluids Struct. 11, 617–626.
(doi:10.1006/jfls.1997.0099)
18. White FM. 2003 Fluid mechanics, 5th edn. New York: McGraw-Hill.
19. Wu J-Z, Lu X-Y, Zhuang L-X. 2007 Integral force acting on a body due to local flow structures.
J. Fluid Mech. 576, 265–286. (doi:10.1017/S0022112006004551)

You might also like