You are on page 1of 9

1

Proceedings of the 37
th
International & 4
th
National Conference on Fluid Mechanics and Fluid Power
FMFP2010
December 16-18, 2010, IIT Madras, Chennai, India
FMFP10 - AM - 01


SPHERE ROLLING DOWN AN INCLINE
SUBMERGED IN A LIQUID

Pravin K. Verekar
Department of Mechanical Engineering
Indian Institute of Science
Bangalore, Karnataka, India
Jaywant H. Arakeri
Department of Mechanical Engineering
Indian Institute of Science
Bangalore, Karnataka, India
pravin@mecheng.iisc.ernet.injaywant@mecheng.iisc.ernet.in

ABSTRACT

A sphere rolling down an inclined plane submerged


in water is studied experimentally in order to
understand the different forces acting on it. This
forms a class of solid-fluid interaction problems that
include sediment transport, movement of gravel on
ocean floor and river bed due to water currents. The
flow development around the rolling sphere is
elucidated in order to highlight its implications on
the nature of hydrodynamic forces that act on the
sphere. Equation of motion for the sphere is solved
numerically and the experimental data is fitted on
these solutions; the best fit gives the values of the
force coefficients.
Keywords: rolling sphere, inclined plane, flow
development, hydrodynamic forces, force coeffi-
cients.

INTRODUCTION

Experiments are done with a sphere rolling


down an inclined plane submerged in quiescent
water. These experiments are conducted to
determine the hydrodynamic force coefficients.
The experimental setup consists of a glass tank
15 cm wide by 14 cm deep by 61 cm long. At
one end, the glass tank is fixed at the base with
two levelling screws on either side, and at the
other end, the tank rests on a spherical pivot
mounted centrally. The levelling screws allow
the tank to be tilted to the desired angle.
The motivation for this study comes from
the need to improve understanding of the solid-
fluid interaction problems such as sediment
transport, movement of gravel on ocean floor
and river bed, etc.


Proceedings of the 37th National & 4th International Conference on Fluid Mechanics and Fluid Power
December 16-18, 2010, IIT Madras, Chennai, India.
FMFP10 - AM - 01
2
EXPERIMENTS

A solid rigid smooth sphere is released from rest on
the inclined bottom glass plane of the tank which is
filled with water. The sphere rolls down the plane
under the influence of gravity. Initial motion of the
sphere is with decreasing acceleration until it attains
terminal velocity. The spheres used for the
experiments along with the inclined plane angles
kept during these experiments are given in Table 1.
The spheres are so chosen because they have good
material homogeneity and are perfectly spherical.
The motion of the sphere is photographed in front
view using Photron FASTCAM PCI R2 model 500
digital camera. Photography for some experimental
runs is done in close view which captures the initial
acceleration portion of the sphere motion while for
other runs it is done in far view which captures the
full motion of the sphere till it attains terminal
velocity. The displacement of the sphere in pixels is
obtained from its translation in the digital images
and the time increment is known from the framing
rate. The conversion scale for distance is based on
the sphere size in the digital image given by pixels
per mm of the sphere diameter. The paired data of
the displacement and time are used later for the
kinematic analyses of the sphere.

Table 1: Experimental settings
Sphere
description
Diameter
(cm)
Specific
gravity,
Inclined plane
angles, u
Ratio of diameter, D
to width of tank, W
1 Acrylic sphere No. 1
2.54

1.18

1.8, 5.7

0.17
2 Acrylic sphere No. 2
5.08

1.19

1.7, 2.8

0.34
3 Pool ball


5.24

1.70

2.9

0.35

The sphere is released using a pair of tongs
made from aluminium wire of diameter 1.5 mm. In
some experiments a pair of tongs made from steel
wire of diameter 1.2 mm is used. To find the angle of
the inclined plane, the difference in height of the
surface of water from the bottom of the tank is
measured at a separation of 50 cm and 30 cm along
the incline of the bottom of the tank; and then the
average is taken of the two values obtained from the
inverse of sine of the ratio of the difference in height
to separation distance.
FLOW DEVELOPMENT AROUND THE
SPHERE

The flow development around the sphere is
explained here. As the sphere begins to roll, the
Reynolds Re based on instantaneous velocity
(hereafter called instantaneous Reynolds
number) are in the creeping viscous flow
regime. For creeping flow or Stokes flow
regime, Re is less than 1; this translates for the
experiments carried out as: (i) for acrylic
sphere No. 1, the instantaneous velocity V of
the sphere less than 0.004 cm/s, and (ii) for
acrylic sphere No. 2 and pool ball as the
velocity less than 0.002 cm/s. What is noticed
is that, the Stokes flow is momentary and
appears to have no consequence on the later
flow development. Creeping viscous flows are
explained in Langlois, 1964.
The starting flow over the rolling sphere
can be treated as irrotational; here the fluid
particles glide over the surface of the sphere
and the velocity field has a potential which is
completely defined by knowing the
instantaneous normal velocity of the surface of
the sphere; this means that the irrotational
motion is entirely without memory (Lighthill,
1986). Cox & Cooker (2000) have found that
for an irrotational flow past a sphere touching a
tangent plane, the flow velocity is singular at
the point of contact and the flow speeds around
the point of contact are large. Also it is noted
that the influence of the tangent plane is small
one radius away and the flow there is very
similar to that for an isolated sphere.
The starting potential flow over the sphere
is brief as the no-slip condition takes hold at
the solid-fluid interface. The fluid particles
next to the surface adhere to it and the
succeeding adjacent fluid layers are sheared
until the outer edge of the boundary layer. The
boundary layer is a highly strained flow and is
the region where all the vorticity is confined.
This thin viscous layer is pressed upon by the
outer potential flow, and its growth is governed
by the momentum diffusivity of the fluid and
3
the convective velocity of the outer flow. The fluid
particles in the boundary layer in overcoming the
viscous friction lose kinetic energy; hence this
retarded layer of flow cannot follow the curvature of
the sphere at the rear side as the fluid particles with
smaller kinetic energy cannot overcome the positive
pressure gradient present in this region and the
boundary layer separates; initially forming a
separation bubble at lower Re, and later in time
shedding lumps of vorticity at higher instantaneous
Re (see Schlitchting, 1968). Boundary conditions no
longer govern the separated flow and as Lighthill
(1986) puts it all of the memory in a fluid flow lies
in its vorticity; which, once generated is subject to
convection and diffusion. Stuart (1963) points out
that at large times, the boundary layer flow
becomes quasi-steady, in the sense that it behaves
like a steady flow of boundary-layer theory with
instantaneous outer flow speed.
Since the flow velocities are higher around the
bottom hemisphere near the point of contact than
those around the top hemisphere, there is a
differential shedding of the vortex sheet (boundary
layer) that rolls up at the rear of the sphere. The
circulating eddies detaching from the lower region
are convected with higher velocity; they are pushed
upwards by the outer potential flow and are pressed
sideways by the slower rotating eddies above them;
rotating fluid elements acquire an upward and
transverse velocity components. This streamwise
spiral vorticity cannot be captured properly in 2-D
flow visualisation.
Adding complexity to the flow features are two
more phenomena peculiar to the rolling sphere: (i) a
primary flow in the boundary layer which is
opposing the potential flow in the top hemisphere
and which is along the potential flow in the lower
hemisphere; and (ii) a secondary flow along the
surface, spiralling away from the poles towards the
equator (Howarth, 1951). Three dimensional
boundary layers are discussed in Moore (1956).
Visualizations of the wake formation behind a
rolling sphere in steady uniform flow have been
done by Stewart et al. (2008) using fluorescein dye.
The Reynolds numbers are varied from 75 to 350.
They have defined a parameter the rotation rate of
the sphere, o which is the ratio of the
tangential velocity on the surface of the sphere
with respect to the centre of the moving sphere
to the translation velocity of the sphere in
ground reference frame. The case o=1
represents sphere rolling down an inclined
plane. Their visualizations show a steady wake
mode for low Re with transition to unsteady
wake mode occurring at around Re=100. In the
steady wake mode, they observe that the
opposing motion of the top surface of the
sphere to the outer flow creates a zone of
recirculating fluid behind the sphere and the
dye escapes this recirculation zone via a single
tail along the centreline of the body. In the
unsteady wake mode, they observe that the
shedding from the top of the sphere takes the
form of hairpin vortices similar to that for an
isolated sphere.
The flow features that evolve in time
around the sphere determine the nature of the
hydrodynamic forces that act on it.

FORCES ACTING ON THE SPHERE

The sphere rolls under the influence of the
component of the resultant of the weight force
minus the buoyancy force along the plane. This
driving force can be expressed as
( )
sin
s f
Vol g u where
s
is the density
of the sphere,
f
is the density of the fluid,
Vol is the volume of the sphere, g is the
acceleration due to gravity and u is the angle
of the incline.
When the driving force accelerates the
sphere in still water, it needs to also accelerate
the surrounding mass of water. This rate of
change of momentum of the surrounding mass
of fluid appears as a resisting force; this is
accounted by considering an increase in the
mass of the sphere by adding a separate mass,
what is called added-mass (also called virtual
mass), to the mass of the sphere. The added-
mass is taken as a factor Ca times the mass of
the displaced fluid. The added-mass force is
4
given by
f
dV
Ca Vol
dt
where Ca is the added-mass
coefficient and V is the velocity of the sphere.
The other hydrodynamic resisting force to the
motion of the sphere, the drag force, originates from
the tangential skin friction (skin drag) and the
difference in the quasi-steady normal pressure at fore
and aft of the sphere (form drag). At the initial times
when the boundary layer has not separated, the
relative value of the skin drag to the total drag is
high; while on separation of the boundary layer, it is
the relative value of the form drag to the total drag
that is higher; the higher form drag corresponds to
the low pressures at the rear of the sphere when the
separation occurs. The drag force is given by
1
2
f
Cd A V V where Cd is the drag coefficient
and A is the projected area of the sphere
perpendicular to the main flow.
Also opposing the motion is the force of
friction at the point of contact. When the sphere is in
pure rolling, this force is given by
2
I dV
R dt
where I
is the mass moment of inertia of the sphere about the
diameter and R is the radius of the sphere.

THEORY

The equation of motion of the sphere is given below.
( )
2
2
1
sin
2
s s f f f
dV dV I dV
Vol gVol Ca Vol Cd V A
dt dt R dt
u =
(1)
The above equation is solved for the solid sphere
where the ratio
s
f

is written as . The following


equation is arrived at.
( ) ( )
2
3
1.4 1 si n
4
dV V
Ca g Cd
dt D
u + =
(2)
During the early part of the motion of the
sphere, when the flow over the sphere is potential,
the dominant hydrodynamic resistance comes from
the inertia of the added-mass. During the later part of
the motion, when the boundary layer has developed
and then separated, the hydrodynamic drag is the
main opposing force.
Similar inclined plane experiments have
been done previously by Carty (1957), Garde
& Sethuram (1969), and Jan & Chen (1997).
The values of Cd measured by Garde &
Sethuram are higher than Cartys values, while
those measured by Jan & Chen are
intermediate. The first two studies are not
concerned with finding Ca, while Jan & Chen
determine Ca as 2.0 using insufficient data.
Chhabra & Ferreira (1999) have solved the
Eq. (1) analytically using Jan & Chens drag
and added-mass coefficient data. They use a
single expression for CdRe relation, which
gives the best fit to the CdRe equations of Jan
& Chen (1997). The CdRe relation used by
them is given below.
5
321.906
Cd 0.861 0.1 Re < 10
Re
= + <
(3)
Equations for the potential flow past a
sphere touching a tangent plane have been
solved numerically by Cox & Cooker (2000).
They find the added-mass coefficient Ca as
0.621.
EXPERIMENTAL OBSERVATIONS AND
ANALYSES

Collateral experiments on unidirectional,
uniform, unsteady flow past the unconstrained
acrylic sphere No. 2 resting on a horizontal
plane in the unsteady water tunnel facility in
the laboratory have shown that the sphere rolls
on the glass surface without slipping. A sphere
rolls without sliding when the rolling friction is
less than the sliding friction. As explained in
Starzhinskii (1982), pure rolling occurs when
the ratio of the coefficient f
r
of rolling friction
to the radius R of the sphere is less than
coefficient f
s
of sliding friction; this statement
is independent of the magnitude of the resultant
driving force when this force passes through
the centroid of the homogeneous sphere.
Practically the coefficient f
r
depends solely on
the materials of the rolling body and of the
inclined plane; and if the materials of the
5
rolling body and of the inclined plane are sufficiently
hard, the force of the rolling friction is very small
(Strelkov, 1978). Usually the ratio f
r
/R is
considerably smaller than the coefficient f
s
of sliding
friction (Starzhinskii, 1982). Verification that the
sphere is in pure rolling is important for the validity
of Eq. (1).
Equation (2) is solved for the velocity V of the
sphere using MATLAB differential equation solver
ode45 (refer Shampine et al., 2003). Cd is taken as
given in Eq. (3) and Ca is treated as a parameter; the
values of Ca are chosen as 0.5, 0.621, 1.0, 1.5, and
2.0. The value 0.5 comes from the potential flow
solution for an isolated sphere; the value 0.621 is
taken from the potential flow solution for a sphere
touching a tangent plane; the value 2.0 is as given by
Jan & Chen (1997) for a rolling sphere on the
incline. The values of , u and D come from the
experimental settings, which are given in Table 1.
The MATLAB code plots following graphs: distance
X travelled by the sphere versus time t, velocity V of
the sphere vs time t, and velocity V of the sphere vs
dimensionless distance X/D. Experimental data
points are superposed on these graphs. Velocity for
the experimental data is obtained by fitting cubic
spline for distance-time points and then
differentiating the spline. Typical graphs for an
experimental run for acrylic sphere No. 1 (diameter
D=2.54cm) and inclination u = 5.7 are shown in
Fig. 1 to Fig. 3, and for acrylic sphere No. 2
(diameter D=5.08cm) and inclination u = 2.8 are
shown in Fig. 6 to Fig. 8.
Figure 1 gives the distance travelled by the
acrylic sphere No. 1 plotted against time for the
entire motion of the sphere till it reaches terminal
velocity. The coloured curves are obtained from the
numerical solutions to the differential equation for
different values of Ca; and it is found that the
experimental points fall close to the green curve (Ca
=0.621) for the entire motion of the sphere. This is
also seen in Fig. 2 and Fig. 3. The sphere attains
terminal velocity at about time t=2.5s and distance
X/D=6. Taking the green curve as the best fit for the
experimental points, the various forces, namely, the
added-mass force, the drag force and the rolling
friction, are calculated as the sphere rolls down the

Fig. 1. Distance vs time diagram

Fig. 2. Velocity vs time diagram

Fig. 3. Velocity vs X/D diagram
6
incline; and the same are normalised by the driving
force. The normalised forces are shown in Fig. 4 and
Fig. 5. It is seen from Fig. 5 that the dominant
influence of the added mass force at starting times is
for a distance X/D up to about 0.3.

Fig. 4. Normalised force vs time diagram

Fig. 5. Normalised force vs X/D diagram

For acrylic sphere No. 2, Fig. 6 gives plot of


distance travelled vs time, Fig. 7 gives plot of
velocity vs time, and Fig. 8 gives plot of velocity vs
X/D. It is seen from these plots that the experimental
points at initial times fall close to the green curve
(Ca =0.621) but later when nearing terminal velocity
(at time t=3.5s and X/D=3) they deviate and fall
below the green line. Calculations show that the Cd
at the terminal velocity for the experimental points is

Fig. 6. Distance vs time diagram

Fig. 7. Velocity vs time diagram

Fig. 8. Velocity vs X/D diagram

7

Fig. 9. Velocity vs time diagram

Fig. 10. Normalised force vs time diagram

Fig. 11. Normalised force vs X/D diagram

1.2 times higher than that given by the green
line. Taking Ca as 0.621 and Cd as 1.2 times
that given by Eq.(3), the differential equation
of the motion of the sphere is solved. This
solution and the experimental points are shown
in Fig. 9. The solution fits the experimental
points adequately, except at the region where
the acceleration ends and the terminal velocity
starts. Taking this solution as the fit, the
different normalised forces are calculated and
shown in Fig. 10 and Fig. 11. From Fig. 11, it
is seen that the dominant influence of the added
mass force at starting times is for a distance
X/D up to about 0.3.
Terminal velocity V
t
is found from the
experimental data. Reynolds number Re
t
is
calculated at terminal velocity. Coefficient of
drag Cd
expt
is calculated by equating drag force
to the driving force, which it balances when the
sphere has attained terminal velocity. Cd
lit.
is
the drag coefficient obtained from Eq.(3),
which is given by Chhabra & Ferreira (1999).
Results are presented in Table 2.
Table 2: Experimental results
Sphere
description
D
cm
D/W u Vt
cm/s
Ret Cdexpt Cdlit.
.
exp
lit
t
Cd
Cd
Acrylic sph. 1 2.54 0.17 1.8 3.9 990 1.23 1.19 1.0
Acrylic sph. 1 2.54 0.17 5.7 7.5 1900 1.06 1.03 1.0
Acrylic sph. 2 5.08 0.34 1.7 5.6 2800 1.19 0.98 1.2
Acrylic sph. 2 5.08 0.34 2.8 7.4 3800 1.13 0.95 1.2
Pool ball 5.24 0.35 2.9 14.9 7800 1.09 0.95 1.2



DISCUSSION AND CONCLUSIONS

It is confirmed experimentally that the Ca for a
sphere rolling down an inclined plane is close
to 0.621 and it holds good at D/W ratio of 0.17
and 0.35. The equation (Eq. (3)) for Cd for the
rolling sphere given by Chhabra and Ferreira
(1999) is valid for D/W ratio 0.17 but not for
0.35. At D/W=0.35, the Cd is higher by 1.2
times that given by the Eq. (3). It may be
mentioned that the Cd-Re relation given by
Eq.(3) is for a rolling sphere down an incline
and cannot be used for a flow past a stationary
8
sphere touching a tangent plane since the wake
profile in the two cases is likely to be different. It
also cannot be used when the rolling of a sphere on a
horizontal plane is induced by a flow of fluid taking
place around it. So the Cd-Re relation for a rolling
sphere is problem specific. No such restriction
applies for Ca except that the sphere is touching a
tangent plane.
From Fig. 5 and Fig. 11, it is seen that the
dominant influence of the added mass force at
starting times is for a distance X/D up to about 0.3.


NOMENCLATURE

u angle of incline
specific gravity

f
density of fluid

s
density of sphere
A projected area of sphere
Ca added-mass coefficient
Cd drag coefficient
Cd
expt
drag coefficient from experiment calculated
at terminal velocity
Cd
lit.
drag coefficient from eq. given by Chhabra
& Ferreira calculated at terminal velocity
D diameter of sphere
f
r
coefficient of rolling friction
f
s
coefficient of sliding friction
g acceleration due to gravity
I mass moment of inertia of sphere about its
diameter
R radius of sphere
Re instantaneous Reynolds number
Re
t
Reynolds number at terminal velocity
t time
V instantaneous velocity of sphere
V
t
terminal velocity of sphere
Vol volume of sphere
W width of tank
X instantaneous distance travelled by sphere

REFERENCES


Chhabra, R. P., Ferreira, J. M., 1999. An
analytical study of the motion of a sphere
rolling down a smooth inclined plane in an
incompressible Newtonian fluid, Powder
Technology, 104 pp. 130138.
Cox, S. J., Cooker, M. J., 2000. Potential flow
past a sphere touching a tangent plane, Journal
of Engineering Mathematics, 38 pp. 335370.
Jan, C., Chen, J., 1997. Movements of a sphere
rolling down an inclined plane, Journal of
Hydraulic Research, 35 pp. 689706.
Langlois, W. E., 1964. Slow Viscous Flow.
Macmillan, New York.
Lighthill, J., 1986. An Informal Introduction to
Theoretical Fluid Mechanics. Clarendon Press,
Oxford.
Moore, F. K., 1956. Three-dimensional
Boundary Layer Theory, Advances in Applied
Mechanics, 4 (2) 159-228.

Schlichting, H., 1968. Boundary Layer Theory,
sixth ed. McGraw-Hill, New York, pp. 2443.
Shampine, L. F., Gladwell, I., Thompson, S.,
2003. Solving ODEs with MATLAB,
Cambridge University Press, New York, pp. 1
131.
Starzhinskii, V. M., 1982. An Advanced
Course of Theoretical Mechanics, Mir Publi-
shers, Moscow, pp. 8889.
Stewart, B. E., Leweke, T., Hourigan, K.,
Thompson, M. C., 2008. Wake formation
behind a rolling sphere, Physics of Fluids, 20,
071704, 14.
Strelkov, S. P., 1978. Mechanics, Mir
Publishers, Moscow, pp. 259269.
9
Stuart, J. T., 1963. Unsteady Boundary Layers. In:
Rosenhead, L., (Ed.), Laminar Boundary Layers.
Clarendon Press, Oxford, pp. 349408.

You might also like