You are on page 1of 3

Simple Model of a Rolling Water-Filled Bottle on an Inclined Ramp

Shihao Lin, Naiwen Hu, Tianchen Yao, Charles Chu, Simona Babb, Jenna Cohen, Giana Sangiovanni,
Summer Watt, Danielle Weisman, James Klep, and Wojciech J. WaleckiEve S. Walecki and Peter S.
Walecki

Citation: Phys. Teach. 53, (2015); doi: 10.1119/1.4935768


View online: http://dx.doi.org/10.1119/1.4935768
View Table of Contents: http://aapt.scitation.org/toc/pte/53/9
Published by the American Association of Physics Teachers
Simple Model of a Rolling Water-Filled
Bottle on an Inclined Ramp
Shihao Lin, Naiwen Hu, Tianchen Yao, Charles Chu, Simona Babb, Jenna Cohen, Giana Sangiovanni,
Summer Watt, Danielle Weisman, James Klep, and Wojciech J. Walecki, North Broward Preparatory School,
Coconut Creek, FL
Eve S. Walecki and Peter S. Walecki, Sunrise Optical LLC, Sunrise, FL

W
e investigate a water-filled bottle rolling down an Similarly for the rolling rigid cylinder we have
incline and ask the following question: is a rolling
bottle better described by a model ignoring all (6)
internal motion where the bottle is approximated by a mate-
rial point sliding down an incline, or is it better described by a where
rigid solid cylinder rolling down the incline without skidding? (7)
The measurements presented here represent a special case of
similar experiments described by K.A. Jackson et al. (see Ref. is the moment of inertia of a solid cylinder having radius r
1 and references within). There exists also a report by Kagan2 and ω is the angular speed of the cylinder, and if no slipping
describing the motion of soda cans rolling on an incline. In occurs the angular speed of the cylinder is given by
our case we investigate motion of the fully filled bottle. We
demonstrate that within accuracy of our experiment the mo- ω = v/r. (8)
tion of the bottle can be described by a simple “frictionless
water” model. The analysis of the dynamics of the bodies Analysis similar to the case of the frictionless sliding point
sliding and rolling on a ramp is a standard component of in- leads to the following formula for acceleration of a rolling
troductory physics classes, and a required component of the cylinder:
Advanced Placement (AP) Physics curriculum.3
(9)

Theory The linear acceleration of the solid cylinder is smaller than
Conservation of mechanical energy for the point sliding that of the sliding point since part of the gravitational poten-
down the incline, having zero initial velocity, implies the fol- tial energy is consumed by kinetic energy of rotation
lowing relationship between kinetic and potential energies:
(10)
(1)
of the rolling cylinder.
where m is the mass of material point, v is its speed, h is height
of the incline, and g is free-fall acceleration. Results
Since the height of the inline is given by In our experiment we used lightweight 2-liter soda bottles,
an inclined ramp at approximately 5° with respect to hori-
h = l sin (α) , (2) zontal direction, a ruler, a digital stopwatch, and a scale. The
weight of the bottle when empty was measured to be
where α is the angle of incline with respect to horizontal di- 0.04 kg and approximately 2.04 kg when fully filled with
rection and l is length of the incline, we find that water. So indeed one can neglect mass of the container. The
water-filled bottle was then released from the top of the ramp.
(3) The time it took the bottle to roll a distance varying from 0.1
m to 0.8 m down the ramp was recorded. Each measurement
It is well known that the acceleration of the objects moving was repeated three times, and the arithmetic average of three
on an incline is constant. The kinematic equation describing measurements was calculated and considered a result of mea-
acceleration of the body starting from rest, surement. The standard deviation of the three measurements
(4) corresponding to the same distance had range between
0.022 s to 0.045 s. We estimate the uncertainty of the position
combined with Eq. (3) leads to the following expression for measurement to be 4 mm.
linear acceleration of the sliding point: The results are presented in Fig. 1. In the same figure we
predicted and plotted a simple “frictionless water” model
a = g sin α. (5) where the internal motion of the liquid can be neglected, and

548 The Physics Teacher ◆ Vol. 53, December 2015 DOI: 10.1119/1.4935768
Conclusion
0.8
Experiment: Liquid water
We have presented a very simple model explaining the
Experiment: Ice rolling motion of a water-filled soda bottle. We have shown
0.6 Theory: Friconless sliding that a model neglecting the coupling between the walls and
the liquid adequately describes results of our experiments.
Theory: Rolling solid
We have shown that water filled bottles travel down inclines
Posion [m]

cylinder
0.4 with acceleration equal to the acceleration of a body sliding
down the same incline in the absence of kinetic friction.

References
0.2
1. K. A. Jackson, J. E. Finick, C. R. Bednarski, and L. R. Clifford
“Viscous and nonviscous models of the partially filled rolling
can,” Am. J. Phys. 64, 277–282 (March 1996).
0 2. David Kagan, “The shaken-soda syndrome,” Phys. Teach. 39,
0 0.5 1 1.5 290–292 (May 2001).
Time [s] 3. The College Board, “Physics course description,” http://
apcentral.collegeboard.com/apc/public/repository/ap-
Fig. 1. Motion of the bottle filled with liquid water (solid dia- physics-course-description.pdf), accessed on Feb. 25, 2012.
monds) and a bottle filled with solid ice (empty squares) slid-
ing down a 5° incline. The solid line describes the “frictionless
sliding” model, where the internal motion of the liquid can
be neglected, while the dashed line describes the “rolling
solid cylinder” model, where the bottle and water are treated
as one rigid cylinder rolling down the incline without sliding.
The typical standard deviation of time measurement for liquid
water measurement corresponding to position on incline
0.50 m was  0.035 s. The position accuracy was estimated to
be 4 mm and was smaller than the symbols used in this graph.

a “rigid water” model where the bottle and water are treated
as one rigid cylinder. Directly from Fig. 1 we see that for the
bottle filled with liquid, the data are clearly more consistent
with the stationary water model. For the bottle filled with ice,
data are more consistent with the rigid cylinder model.
From the data presented in Fig. 1 we see that the friction-
less water model is indeed a much more adequate description
of the motion of the bottle.
The results are consistent with results reported by Kagan.2
The coupling between the walls and liquid is more important
in small-diameter containers used by Kagan rather than in
large-diameter ones that we used. This fact can easily be seen
as a consequence of Newton’s second law for rotating bodies.
The torque of viscous forces acting on the liquid is propor-
tional to the area of the side wall of the container and its ra-
dius or  Lr2, where L is the length of the bottle. The moment
of inertia of the spinning cylinder is proportional to the mass
and the square of the radius or  Lr4. Therefore, we expect
that the angular acceleration of the spinning liquid should be
proportional to the inverse of the square of the radius or α 
r–2. Since a soda can diameter is approximately 6.5 cm, and
the diameter of a 2-liter soda bottle is approximately 10.8 cm,
we expect that coupling is less important in the case of the
large 2-liter bottle than in the smaller soda can.

The Physics Teacher ◆ Vol. 53, December 2015 549

You might also like