A Coriolis Tutorial: Part 1: The Coriolis Force, Inertial and Geostrophic Motion

Download as pdf or txt
Download as pdf or txt
You are on page 1of 57

a Coriolis tutorial

Part 1: the Coriolis force, inertial and geostrophic motion

James F. Price
Woods Hole Oceanographic Institution
Woods Hole, Massachusetts, 02543
www.whoi.edu/science/PO/people/jprice [email protected]

Version 6 May 30, 2013

Abstract: This essay is the first of a three-part introduction to the Coriolis force and its consequences
for the atmosphere and ocean. It is intended for students who are beginning a quantitative study of
geophysical fluid dynamics and who have some background in classical mechanics and applied
mathematics.

1
The equation of motion appropriate to a steadily rotating reference frame includes two terms that
account for accelerations that arise from the rotation of the reference frame, a centrifugal force and a
Coriolis force. In the case of an Earth-attached reference frame, the centrifugal force is effectively
subsumed into the gravity field. The Coriolis force has a very simple mathematical form, −2Ω Ω ×V0 M,
where Ω is Earth’s rotation vector, V0 is the velocity observed from the rotating frame and M is the
parcel mass. The Coriolis force is perpendicular to the velocity and so tends to change velocity
direction, but not velocity amplitude, i.e., the Coriolis force does no work. It nevertheless has a
profound effect on large scale, low frequency currents and winds.
If the Coriolis force is the only force acting on a moving parcel, then the velocity vector of the
parcel will be turned anti-cyclonically (clockwise in the northern hemisphere) at the rate − f , where
f = 2Ωsin(latitude) is the important Coriolis parameter. These free motions, often termed inertial
oscillations, are a first approximation of the upper ocean currents generated by a transient wind event. If
the Coriolis force is balanced by a steady force, say a pressure gradient, then the resulting steady wind
or current will be in a direction that is perpendicular to the pressure gradient. An approximate
geostrophic balance of this kind is the defining characteristic of the large scale, low frequency
circulation of the atmosphere and oceans outside of the tropics.

Cover graphic. Earth’s rotation vector, Ω , maintains a nearly steady bearing close to Polaris, commonly
called the Pole Star or North Star. Earth thus has a specific orientation with respect to the universe at large.
This is manifest as a marked anisotropy of most large-scale circulation phenomena, e.g., the east-west
asymmetry of ocean gyres and the westward propagation of low frequency waves and eddies.

Contents
1 Large-scale flows of the atmosphere and ocean 3
1.1 Models and reference frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.1 Classical mechanics observed from an inertial reference frame . . . . . . . . . . 7
1.1.2 Classical mechanics observed from a rotating, noninertial reference frame . . . . 8
1.2 The goal and the plan of this essay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 About this essay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2 Noninertial reference frames 11


2.1 Kinematics of a linearly accelerating reference frame . . . . . . . . . . . . . . . . . . . 11
2.2 Kinematics of a rotating reference frame . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.1 Transforming the position, velocity and acceleration vectors . . . . . . . . . . . 14

2
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 3

2.2.2 Stationary ⇒ Inertial; Rotating ⇒ Earth-Attached . . . . . . . . . . . . . . . . 20


2.2.3 Remarks on the transformed equation of motion . . . . . . . . . . . . . . . . . . 22

3 Inertial and noninertial descriptions of elementary motions 24


3.1 Switching sides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2 To get a feel for the Coriolis force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2.1 Zero relative velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.2.2 With relative velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 An elementary projectile problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3.1 From the inertial frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3.2 From the rotating frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4 Appendix to Section 3: Cylindrical polar coordinates. . . . . . . . . . . . . . . . . . . . 33

4 A reference frame attached to the rotating Earth 35


4.1 Cancellation of the centrifugal force . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.1 Earth’s (slightly chubby) figure . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.2 Vertical and level in an accelerating reference frame . . . . . . . . . . . . . . . 37
4.1.3 The equation of motion for an Earth-attached frame . . . . . . . . . . . . . . . . 38
4.2 Coriolis force on motions in a thin, spherical shell . . . . . . . . . . . . . . . . . . . . . 39
4.3 Why do we insist on the rotating frame equations? . . . . . . . . . . . . . . . . . . . . 41
4.3.1 Inertial oscillations from an inertial frame . . . . . . . . . . . . . . . . . . . . . 41
4.3.2 Inertial oscillations from the rotating frame . . . . . . . . . . . . . . . . . . . . 44

5 A dense parcel on a rotating slope 46


5.1 Inertial and geostrophic motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.2 Energy balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

6 Summary 54

7 Supplementary material 57

1 Large-scale flows of the atmosphere and ocean

The large-scale, low frequency flows of Earth’s atmosphere and ocean take the form of circulations
around centers of high or low pressure. Global-scale circulations include the broad belt of westerly
wind that encircles the mid-latitudes in both hemispheres (Fig. 1) and gyres that fill ocean basins (Fig.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 4

2). Smaller scale circulations often dominate the weather. Hurricanes and mid-latitude storms have a
more or less circular flow around a low pressure center, and many regions of the ocean are filled with
slowly revolving eddies having a diameter of several hundred kilometers. The pressure anomaly that is
associated with each of these circulations can be understood as the direct consequence of a mass excess
or deficit (high or low hydrostatic pressure) in the overlying fluid.
What is at first surprising is that large scale mass anomalies of the kind seen in Figs. 1 and 2 persist
for many days or weeks even in the absence of an external momentum or energy source. The flow of
mass that would be expected to accelerate down the pressure gradient and disperse the associated mass
anomaly does not occur. Instead, large-scale, low frequency winds and currents are observed to flow in
a direction that is almost parallel to lines of constant pressure; the sense of the flow is clockwise around
high pressure centers (northern hemisphere) and anti-clockwise around low pressure centers. The
direction of flow is reversed in the southern hemisphere. From this we can infer that the pressure
gradient force, which is normal to lines of constant pressure, must be balanced approximately by a
second force that acts to deflect horizontal winds and currents to the right of the velocity vector in the
northern hemisphere and to the left of the velocity vector in the southern hemisphere. This deflecting
force is the Coriolis force1 and is the topic of this essay.(Occasional footnotes provide references,
background knowledge and homework assignments. They may be skipped on first reading.) A balance
between a pressure gradient force and the Coriolis force is called a geostrophic balance, and an
approximate or quasi- geostrophic balance is the defining characteristic of large scale, low frequency
(for now, periods greater than one day) atmospheric and oceanic flows outside of equatorial regions.2
We attribute profound physical consequences to the Coriolis force, and yet cannot point to a
physical interaction as the cause of the Coriolis force in the straightforward way that hydrostatic
pressure anomalies are related to the mass field. Rather, the Coriolis force arises from motion itself,
combined with the necessity that we observe the atmosphere and ocean from an Earth-attached and thus
rotating, noninertial reference frame. In this respect the Coriolis force is quite different from other
important forces acting on geophysical fluids (pressure gradients, friction) in ways and with
consequences that are discussed throughout this essay.
1 After the French physicist and engineer, Gaspard-Gustave de Coriolis, 1792-1843, whose seminal contributions include
the systematic derivation of the rotating frame equation of motion and the development of the gyroscope. An informative
history of the Coriolis force is by A. Persson, ’How do we understand the Coriolis force?’, Bull. Am. Met. Soc., 79(7),
1373-1385 (1998).
2 You must be wondering . . . ’What’s it do right on the equator?’ (S. Adams, It’s Obvious You Won’t Survive by Your

Wits Alone, p. 107, Andrews and McNeil Press, Kansas City, Kansas, 1995). The horizontal component of the Coriolis
force must vanish right on the equator (by symmetry) and hence the contrast between mid-latitude and equatorial phenom-
ena, and specifically the velocity and pressure relationships, is something of great interest here. To start, how would you
characterize the (near-) equatorial wind and pressure relationship of Fig. 1? This one chart is not particularly revealing,
and so it will be helpful to take a look also at the 500 hPa charts (heights, winds and temperatures) available online at
http://www.nrlmry.navy.mil/metoc/nogaps/
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 5

Figure 1: A weather map showing conditions at 500 mb over the North Atlantic on 20 March, 2004,
produced by the Fleet Numerical Meteorology and Oceanography Center (500 mb is a middle level of
the atmosphere). Variables are temperature (colors, scale at right is ◦C), the height of the 500 mb pressure
surface (white contours, values are meters above sea level) and the wind vector (as ’barbs’ at the rear of
the vector, one thin barb = 10 knots ≈ 5 m s−1 , one heavy barb = 50 knots). Several important phenomena
are evident on this map. (1) The winds at mid-latitudes were mainly westerly, i.e., winds that blow from
the west toward the east. The broad band of westerly winds often includes one or several maxima, or
jet stream(s), here between about 40◦ to 50◦ N. (2) Within the westerly wind band, the 500 mb surface
sloped downward toward higher latitude, roughly 500 m per 1000 km where wind speed was roughly
U ≈ 40 m s−1 . There was thus a small, but significant component of gravity that was parallel to the 500
mb surface, −g∇η , and directed from south to north. (3) Wind vectors appear to be nearly parallel to
the contours of constant height everywhere poleward of about 10◦ latitude, and greater wind speed was
coincident with stronger slope. The wind and pressure fields were thus in an approximate geostrophic
balance, for the east-west wind component, f U ≈ g∂ η /∂ y, where f is the Coriolis parameter and f U is
the Coriolis force (per unit mass), the topic of Part 1 of this essay. (4) Any particular realization of the
jet stream is likely to exhibit wave-like undulations called westerly waves that are a key phenomenon
of mid-latitude weather and climate. The longest east-west wavelengths (the example evident here has
a wavelength of roughly 9000 km) often appear to be almost stationary with respect to the Earth, and
so must be propagating westward with respect to the spatially-averaged westerly wind. Westerly waves
having a shorther wavelength (a few thousand kilometers) are generally swept downstream with the
westerlies.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 6

Aviso, SSH, mean of 2007


o
60 N

2
o
45 N

1.5
30oN

1
o
15 N

0.5
0o

100oW 80oW 60oW 40oW 20oW 0o

Figure 2: The 2007 annual mean of sea surface height (SSH). These remarkable data are thanks to
the Aviso project (with support from Cnes, http://www.aviso.oceanobs.com/duacs/). The color scale at
right is in meters. The region depicted here is comparable to that of Fig.1, and the field has a very
similar meaning; SSH is effectively a constant pressure, frictionless surface that is displaced slightly but
significantly from level. The change in SSH is only about 2 m from a low in the western subpolar gyre
(55◦ N and 50◦ W) to a high in the western subtropical gyre and Caribbean Sea (25◦ N and 70◦ W).
Note that by far the largest gradient of SSH is found just inside the western boundary of the subtropical
gyre; the gradient is much smaller over the interior of the ocean basins. What keeps SSH displaced away
from level? We do not have direct measurement of ocean currents at the same spatial resolution (though
surface drifter measurements come close, see http://www.aoml.noaa.gov/phod/dac/gdp science.php —
you may have to type this into your web browser), but nevertheless we can be confident that the horizontal
gravitational force associated with this tilted SSH is nearly balanced by the Coriolis force acting upon
horizontal currents, i.e., we infer a geostrophic momentum balance, just as occurs in the westerly winds
of Fig. 1.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 7

1.1 Models and reference frames

This essay proceeds inductively, developing and adding new concepts one by one rather than deriving
them from a comprehensive starting point. In that spirit, the first physical model considered here in Part
1 will be a single, isolated fluid particle, or ’parcel’. This is a drastic and for most purposes untenable
idealization of a fluid. Winds and currents, like all macroscopic fluid flows, are effectively a continuum
of parcels that interact in three-dimensions; the motion of any one parcel is connected by pressure
gradients and by friction (diffusion of momentum) to the motion of essentially all of the other parcels
that make up the flow. This global dependence is at the very heart of fluid mechanics, but can be set
aside here because the Coriolis force on a given parcel depends only upon the velocity of that parcel. By
starting with this highly simplified model we can develop intuition for the Coriolis force that will carry
over intact to more realistic and much more complex fluid models, e.g., the shallow water model of
Parts 2 and 3. What will go missing in this single parcel model is that the external forces on a parcel (the
F below) must be prescribed in a way that can take no account of global dependence. For that reason
the phenomena that arise in a single parcel model are quite limited, but are nevertheless a recognizable
subset of the phenomena that arise in more realistic fluid models and in the real atmosphere and ocean.

1.1.1 Classical mechanics observed from an inertial reference frame

If the parcel is observed from an inertial reference frame3 then the classical (Newtonian) equation of
motion is just
d(MV)
= F + g∗ M,
dt
where d/dt is an ordinary time derivative, V is the velocity in a three-dimensional space, and M is the
parcel’s mass. The parcel mass (or fluid density) will be presumed constant in all that follows, and the
equation of motion rewritten as
dV
M = F + g∗M, (1)
dt
where F is the sum of the forces that we can specify a priori given the complete knowledge of the
environment, e.g., frictional drag with the sea floor, and g∗ is gravitational mass attraction. These are
3 ’Inertia’has Latin roots in+artis meaning without art or skill and secondarily, resistant to change. Since Newton’s
Principia physics usage has emphasized the latter: a parcel having inertia will remain at rest, or if in motion, continue
without change unless subjected to an external force. A ’reference frame’ is comprised of a coordinate system that serves
to arithmetize the position of parcels, a clock to tell the time, and an observer who makes an objective record of positions
and times. A reference frame may or may not be attached to a physical object. In this essay we suppose purely classical
physics so that measurements of length and of time are identical in all reference frames. This common sense view of space
and time begins to fail when velocities approach the speed of light, which is not an issue here. An inertial reference frame is
one in which all parcels have the property of inertia and in which the total momentum is conserved, i.e., all forces occur as
action-reaction force pairs. How this plays out in the presence of gravity will be discussed briefly in Section 3.1.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 8

said to be central forces insofar as they act in a radial direction between parcels, or in the case of
gravitational mass attraction, between parcels and the Earth.4
This inertial frame equation of motion has two fundamental properties that are noted here because
we are about to give them up:
Global conservation. For each of the central forces acting on the parcel there will be a corresponding
reaction force acting on the environment that sets up the force. Thus the global time rate of change of
momentum (global means parcel plus the environment) due to the sum of all of the central forces
F + g∗ M is zero, and so the global momentum is conserved. Usually our attention is focused on the local
problem, i.e., the parcel only, with this global conservation taken for granted and not analyzed explicitly.
Invariance to Galilean transformation. Eqn. (1) should be invariant to a steady, linear translation of
the reference frame, often called a Galilean transformation, because only relative motion has physical
significance. Thus a constant velocity added to V will cause no change in the time derivative, and
should as well cause no change in the forces F or g∗M. Like the global balance just noted, this
fundamental property is not invoked frequently, but is a powerful guide to the form of the forces F. For
example, a frictional force that satisfies Galilean invariance should depend upon the difference of the
parcel velocity with respect to a surface or adjacent parcels, and not the parcel velocity only.

1.1.2 Classical mechanics observed from a rotating, noninertial reference frame

When it comes to the analysis of the atmosphere or ocean we always use a reference frame that is
attached to the rotating Earth — true (literal) inertial reference frames are not accessible to observation
and wouldn’t be desirable even if they were. Some of the reasons for this are discussed in a later section,
4.3, but for now we are concerned with the consequence that, because of the Earth’s rotation, an
Earth-attached reference frame is significantly noninertial for the large-scale, low-frequency motions of
the atmosphere and ocean. Eqn. (1) does not hold for such motions, even as a first approximation. The
equation of motion appropriate to an Earth-attached, rotating reference frame (examined in detail in
Sections 2 and 4.1) is instead
dV0
M = − 2Ω Ω ×V0 M + F0 + gM, (2)
dt
where the prime on a vector indicates that it is observed from the rotating frame, Ω is Earth’s rotation
vector, gM is the time-independent inertial force, gravitational mass attraction plus the centrifugal force
associated with Earth’s rotation and called simply ’gravity’ (discussed further in Section 4.1). Our
Ω ×V0 M, commonly called the Coriolis force in geophysics. The Coriolis force
interest is the term, − 2Ω
4 Unless it is noted otherwise, the acceleration that is observable in a given reference frame will be written on the left-hand
side of an equation of motion, as in Eqn. (1), even when the acceleration is considered to be the known quantity. The forces,
i.e., everything else, will be written be on the right-hand side of the equation. The parcel mass M is not considered variable
here, and M may be divided out, leaving all terms with physical dimensions [length time−2 ], i.e., accelerations. Even then,
the left and right-hand side term(s) will be called ’acceleration’ and ’force(s)’.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 9

has a very simple mathematical form; it is always perpendicular to the parcel velocity and will thus act
to deflect the velocity unless it is balanced by a quasi-steady F0 , often a pressure gradient as noted in the
opening paragraph and in the discussion of Figs. 1 and 2.

1.2 The goal and the plan of this essay

Eqn. (2) applied to geophysical flows is not controversial and if our intentions were strictly practical we
could accept the Coriolis force as given and get on with the applications. In this essay we take the time
to consider What is the Coriolis force? A tour of the classical mechanics literature yields a variety
of answers, that it is a pseudo force, a virtual force, an apparent force, and most equivocal of all, a
fictitious correction force.5 This is not the clear answer we had hoped for, and may now begin to
wonder — is the Coriolis force something ’real’ or is it just a mathematical device to make things come
out right? To answer this will require a rather slow and careful journey from Eqn. (1) to Eqn. (2) so that
at the end we will understand and so be able to explain6 the origin, the basic properties and some of the
consequences of the Coriolis force.
We have already noted that the Coriolis force is an inertial force (reviewed in Section 2.1) that
arises from the rotation of a reference frame. The origin of the Coriolis force is thus a matter of
kinematics, i.e., mathematics rather than physics (Section 2.2), which is at least partially why the
Coriolis force can be hard to grasp. The best word or words to label the Coriolis term is less clear than
is Eqn. (2); the choice made here is just plain ’Coriolis force’ on the basis of what the Coriolis term
does to the observed motion of the atmosphere and ocean, considered here in Sections 3 and 5, and in
considerably more depth in Parts 2 and 3.
Two very simple applications of the rotating frame equation of motion are considered in Section 3.
These illustrate the often marked difference between inertial and rotating frame descriptions of the same
phenomenon, and they also show that the rotating frame equation of motion (2) does not retain the
fundamental properties of the inertial frame Eqn. (1) noted above. Eqn. (2) applies on a rotating Earth
or a planet, where the centrifugal force associated with planetary rotation is canceled (Section 4). The
rotating frame equation of motion thus treats only the comparatively small relative velocity, i.e., winds
5
The latter is by by J. D. Marion, Classical Mechanics of Particles and Systems (Academic Press, NY, 1965), who de-
scribes the plight of a rotating observer as follows (the double quotes are his): ‘... the observer must postulate an additional
force - the centrifugal force. But the ”requirement” is an artificial one; it arises solely from an attempt to extend the form of
Newton’s equations to a non inertial system and this may be done only by introducing a fictitious ”correction force”. The same
comments apply for the Coriolis force; this ”force” arises when attempt is made to describe motion relative to the rotating
body.’ Rotating observers do indeed have to contend with inertial forces that are not found in otherwise comparable inertial
frames, but these inertial forces are not ad hoc corrections as Marion’s quote (here taken out of context) seems to imply.
6 ’Explanation is indeed a virtue; but still, less a virtue than an anthropocentric pleasure.’ B. van Frassen, ’The pragmatics

of explanation’, in The Philosophy of Science, Ed. by R. Boyd, P. Gasper and J. D. Trout. (The MIT Press, Cambridge
Ma, 1999). This pleasure of understanding is the true goal of this essay, but clearly the Coriolis force has great practical
significance for the atmosphere and ocean and for those of us who study their motions.
1 LARGE-SCALE FLOWS OF THE ATMOSPHERE AND OCEAN 10

and currents. This is a great advantage compared with the inertial frame equation of motion which has
to treat all of the motion, and the resulting gain in simplicity more than compensates for the admittedly
peculiar properties of the Coriolis force. Sections 4 and 5 will discuss two kinds of motion that are
directly dependent upon the Coriolis force, a free oscillation, often called an inertial oscillation, and a
forced and possibly steady motion, usually called a geostrophic wind or current when the force F0 is a
pressure gradient.7

1.3 About this essay

This essay has been written for students who are beginning a quantitative study of geophysical fluid
dynamics and who have some background in classical mechanics and applied mathematics. Rotating
reference frames and the Coriolis force will very likely have been encountered before coming here,
either within a classical mechanics course or in a fluid mechanics course or textbook that treats
geophysical flows.8 There is nothing fundamental and new regarding the Coriolis force that could
possibly be added here, but the hope is that this essay will make a useful supplement by providing
greater mathematical detail than is possible in most fluid dynamics texts, and by emphasizing
geophysical phenomena that are missed or outright misconstrued in most classical mechanics texts.9,10
Geophysical fluid dynamics is all about fluids in motion, and the electronic version of this essay
includes links to animations and to source codes of numerical models that provide a much more vivid
7 All this talk of ’forces, forces, forces’ is already becoming tedious and seems a little archaic. Modern dynamics is more
likely to be developed around the concepts of energy, action and minimization principles, which are very useful in some
special classes of fluid flow. However, it remains that the majority of fluid mechanics proceeds along the path of Eqn. (1) laid
down by Newton. In part this is because mechanical energy is not conserved in most real fluid flows and in part because the
interaction between a fluid parcel and its surroundings is often at issue, friction for example, and is usually best-described in
terms of forces.
8 Classical mechanics texts in order of increasing level: A. P. French, Newtonian Mechanics (W. W. Norton Co., 1971); A.

L. Fetter and J. D. Walecka, Theoretical Mechanics of Particles and Continua (McGraw-Hill, NY, 1990); C. Lanczos, The
Variational Principles of Mechanics (Dover Pub., NY, 1949). Textbooks on geophysical fluid dynamics emphasize mainly the
consequences of Earth’s rotation; excellent introductions at about the level of this essay are by J. R. Holton, An Introduction to
Dynamic Meteorology, 3rd Ed. (Academic Press, San Diego, 1992), and by B. Cushman-Roisin, Introduction to Geophysical
Fluid Dynamics (Prentice Hall, Engelwood Cliffs, New Jersey, 1994). Somewhat more advanced and highly recommended
for the topic of geostrophic adjustment is A. E. Gill, Atmosphere-Ocean Dynamics (Academic Press, NY, 1982) and for waves
generally, J. Pedlosky, Waves in the Ocean and Atmosphere, (Springer, 2003).
9 There are several essays or articles that, like this one, aim to clarify the Coriolis force. A fine treatment in great depth

is by H. M. Stommel and D. W. Moore, An Introduction to the Coriolis Force (Columbia Univ. Press, 1989); the present
Section 4.1 owes a great deal to their work. A detailed analysis of particle motion including the still unresolved matter
of the apparent southerly deflection of dropped particles is by M. S. Tiersten and H. Soodak, ‘Dropped objects and other
motions relative to a noninertial earth’, Am. J. Phys., 68(2), 129–142 (2000). A good web page for general science students
is http://www.ems.psu.edu/%7Efraser/Bad/BadFAQ/BadCoriolisFAQ.html
10 The Coriolis force also has engineering applications; it is exploited to measure the angular velocity required for vehicle

control systems, http://www.siliconsensing.com, and to measure mass transport in fluid flow, http://www.micromotion.com.
2 NONINERTIAL REFERENCE FRAMES 11

depiction of these motions than is possible in a hardcopy.


This essay, along with Parts 2 and 3 and all associated materials, may be freely copied and
distributed for any educational purpose. They may be cited by the MIT Open Course Ware address.11
The first version of this essay was released to the internet in 2003, and since then the text and models
have been revised and expanded several times. The most up-to-date version may be downloaded via the
link in Section 7. Comments and questions are encouraged and may be sent directly to the author at
[email protected]
Financial support during the preparation of these essays was provided by the Academic Programs
Office of the Woods Hole Oceanographic Institution. Additional salary support has been provided by
the U.S. Office of Naval Research. Terry McKee of WHOI is thanked for her expert assistance with
Aviso data. Tom Farrar of WHOI, Pedro de la Torre of KAUST, Adam Laux of Siena Italy, Ru Chen of
MIT/WHOI, Peter Gaube of OSU/COAS, Jennifer Van Wakeman of OSU/COAS and Iam-Fei Pun of
WHOI are all thanked for carefully proof-reading a draft of this essay. Jiayan Yang, Xin Huang and
Dennis McGillicuddy of WHOI are thanked for their insightful comments and suggestions on Part 3.

2 Noninertial reference frames

The first step toward understanding the origin of the Coriolis force is to describe the origin of inertial
forces in the simplest possible context, a pair of reference frames that are represented by displaced
coordinate axes, Fig. (3). Frame one is labeled X and Z and frame two is labeled X 0 and Z 0 . It is helpful
to assume that frame one is stationary and that frame two is displaced relative to frame one YHby a
time-dependent vector, Xo (t). The measurements of position, velocity, etc. of a given parcel will thus
be different in frame two vs. frame one. Just how the measurements differ is a matter of kinematics;
there is no physics involved until we define the acceleration of frame two and use the accelerations to
write an equation of motion, e.g., Eqn. (2).

2.1 Kinematics of a linearly accelerating reference frame

If the position vector of a given parcel is X when observed from frame one, then from within frame two
the same parcel will be observed at the position

X0 = X − Xo .
11 Price,James F., 12.808 Supplemental Material, Topics in Fluid Dynamics: Dimensional Analysis, the Coriolis Force,
and Lagrangian and Eulerian Representations, http://ocw.mit.edu/ans7870/resources/price/index.htm (date accessed) License:
Creative commons BY-NC-SA.
2 NONINERTIAL REFERENCE FRAMES 12

1.8

1.6

2 1.4
Figure 3: Two reference frames
1.8 1.2 are represented by coordinate axes
1.6 Z’ 1 that are displaced by the vector Xo
1.4 0.8 that is time-dependent. In this Sec-
1.2 0.6 tion 2.1 we consider only a rela-
Z 1 0.4
X’ tive translation, so that frame two
X
0.8 0.2
maintains a fixed orientation with
0.6 0
0 0.5 1 1.5 2 respect to frame one. The rotation
0.4
X‘ of frame two will be considered be-
0.2 Xo
0
ginning in Section 2.2.
0 0.5 1 1.5 2
X

The position vector of a parcel thus depends upon the reference frame. Suppose that frame two is
translated and possibly accelerated with respect to frame one, while maintaining a constant orientation
(rotation will be considered shortly). If the velocity of a parcel observed in frame one is dX/dt, then in
frame two the same parcel will be observed to have velocity

dX0 dX dXo
= − .
dt dt dt
The accelerations are similarly d 2X/dt 2 and

d 2 X0 d 2 X d 2 Xo
= 2 − . (3)
dt 2 dt dt 2
We are going to assume that frame one is an inertial reference frame, i.e., that parcels observed in frame
one have the property of inertia so that their momentum changes only in response to a force, F, i.e., Eqn.
(1). From Eqn. (1) and from Eqn. (3) we can easily write down the equation of motion for the parcel as
it would be observed from frame two:
d 2 X0 d 2 Xo
M = − M + F + g∗M. (4)
dt 2 dt 2

Terms of the sort −(d 2Xo /dt 2)M appearing in the frame two equation of motion (4) will be called
’inertial forces’, and when these terms are nonzero, frame two is said to be ’noninertial’. As an example,
suppose that frame two is subject to a constant acceleration, d 2 Xo/dt 2 = A that is upward and to the
right in Fig. (3). From Eqn. (4) it is evident that all parcels observed from within frame two would then
appear to accelerate with a magnitude and direction −A, downward and to the left, and which is, of
course, exactly opposite the acceleration of frame two with respect to frame one. An inertial force
results when we multiply this acceleration by the mass of the parcel. Thus an inertial force is exactly
proportional to the mass of the parcel, regardless of what the mass is. But clearly, the origin of the
2 NONINERTIAL REFERENCE FRAMES 13

inertial force is the acceleration, −A, imposed by the accelerating reference frame, and not a force per
se. Inertial forces are in this respect indistinguishable from gravitational mass attraction which also has
this property. If an inertial force is dependent only upon position, as is the centrifugal force due to
Earth’s rotation (Section 4.1), as is gravitational mass attraction, then it might as well be added with g∗
to denote a single, time-independent acceleration field, usually termed gravity and denoted by g. In fact,
this combined mass attraction plus centrifugal acceleration is the acceleration field that may be
observed directly, for example by a plumb line.12 But, unlike gravitational mass attraction, there is no
interaction between particles involved in an inertial force, and hence there is no action-reaction force
pair associated with an inertial force. Global momentum conservation thus does not obtain in the
presence of inertial forces. There is indeed something equivocal about the phenomenon we are calling
an inertial force, and it is not unwarranted that some authors have deemed them to be ’virtual’ or
’fictitious correction’ forces.5
Whether an inertial force is problematic or not depends entirely upon whether d 2Xo /dt 2 is known
or not. If it should happen that the acceleration of frame two is not known, then all bets are off. For
example, imagine observing the motion of a plumb bob within an enclosed trailer that was moving
along in irregular, stop-and-go traffic. The bob would be observed to lurch forward and backward
unexpectedly, and we would soon conclude that studying dynamics in such an uncontrolled, noninertial
reference frame was going to be a very difficult endeavor. Inertial forces could be blamed if it was
observed that all of the physical objects in the trailer, observers included, experienced exactly the same
unaccounted acceleration. In many cases the relevant inertial forces are known well enough to use
noninertial reference frames with great precision, e.g., the topography of Earth’s gravity field must be
known to within a few tens of centimeters to interpret sea surface altimetry data of the kind seen in Fig.
(2)13 and the Coriolis force can be readily calculated as in Eqn. (2) knowing only Earth’s rotation vector
and the parcel velocity.
In the specific example of a translating reference frame considered here (Fig. 3) one could just as
well transform the observations made from frame two back into the inertial frame one, use the inertial
frame equation of motion to make a calculation, and then transform back to frame two if required. By
that tactic we could avoid altogether the seeming delusion of an inertial force. However, when it comes
to the observation and analysis of Earth’s atmosphere and ocean, there is really no choice but to use an
Earth-attached and thus rotating and noninertial reference (discussed in Section 4.3). That being so, we
have to contend with the Coriolis force, an inertial force that arises from the rotation of an
12 A plumb bob is nothing more than a weight, the bob, that hangs from a string, the plumb line (and plumbum is the Latin
for lead, Pb). When a plumb bob is at rest in a given reference frame, the plumb line is parallel to the local acceleration field
of that reference frame. If the bob is displaced and released, it will oscillate as a simple pendulum. The observed period of
small amplitude oscillations, P, can be used to infer the magnitude of the acceleration, g = L/(P/2π )2 , where L is the length
of the plumb line. If the reference frame is attached to the rotating Earth, then the motion of the bob will be effected also by
the Coriolis force, in which case the device is often termed a Foucault pendulum, discussed further in a later footnote 30.
13 Earth’s gravity field is the object of extensive and ongoing survey by some of the most elegant instruments ever flown in

space, see http://www.csr.utexas.edu/grace/ and http://www.esa.int/Our Activities/Operations/GOCE operations


2 NONINERTIAL REFERENCE FRAMES 14

Earth-attached frame. The kinematics of rotation add a small complication that is taken up in the next
Section 2.2. But if you followed the development of Eqn. (4), then you already understand the origin of
inertial forces, including the Coriolis force.

2.2 Kinematics of a rotating reference frame

The equivalent of Eqn. (4) for the case of a steadily rotating reference frame is necessary to reveal the
Coriolis force. Reference frame one will again be assumed to be stationary and defined by a triad of
orthogonal unit vectors, e1 , e2 and e3 (Fig. 4). A parcel P can then be located by a position vector X
X = e1 x1 + e2x2 + e3 x3 , (5)
where the Cartesian (rectangular) components, xi , are the projection of X onto each of the unit vectors
in turn. It is useful to rewrite Eqn. (5) using matrix notation. The unit vectors are made the elements of
a row matrix,
E = [e1 e2 e3 ], (6)
and the components xi are taken to be the elements of a column matrix,
 
x1
X =  x2  . (7)
x3
Eqn. (5) may then be written in a way that conforms with the usual matrix multiplication rules as
X = EX. (8)

The vector X and its time derivatives are presumed to have an objective existence, i.e., they
represent something physical that is unaffected by our arbitrary choice of a reference frame.
Nevertheless, the way these vectors appear clearly does depend upon the reference frame (Fig. 4) and
for our purpose it is essential to know how the position, velocity and acceleration vectors will appear
when they are observed from a steadily rotating reference frame. In a later part of this section we will
identify the rotating reference frame as an Earth-attached reference frame and the stationary frame as
one aligned on the distant fixed stars. It is assumed that the motion of the rotating frame can be
represented by a time-independent rotation vector, Ω . The e3 unit vector can be aligned with Ω with no
loss of generality, Fig. (4a). We can go a step further and align the origins of the stationary and rotating
reference frames because the Coriolis force is independent of position (Section 2.2).

2.2.1 Transforming the position, velocity and acceleration vectors

Position: Back to the question at hand: how does this position vector look when viewed from a
second reference frame that is rotated through an angle θ with respect to the first frame? The answer is
2 NONINERTIAL REFERENCE FRAMES 15

e
1 Ω ’e2 2

0.8

0.6
e3, ’e3
P x2
0.4

θ
0.2 L2 ’e1
X x’
e X 1
0 2

1 ’e2 ’e
1
θ x2’ L1
e
0.5 θ 1
θ
b x1 e1
0
a 1
0.5
−0.5 0
−0.5
stationary ref frame

Figure 4: (a) A parcel P is located by the tip of a position vector, X. The stationary reference frame
has solid unit vectors that are presumed to be time-independent, and a second, rotated reference frame
has dashed unit vectors that are labeled `ei . The reference frames have a common origin, and rotation
is about the e3 axis. The unit vector e3 is thus unchanged by this rotation and so `e3 = e3. This holds
also for Ω 0 = Ω , and so we will use Ω exclusively. The angle θ is counted positive when the rotation is
counterclockwise. (b) The components of X in the stationary reference frame are x1 , x2, x3 , and in the
rotated reference frame they are x01 , x02, x03.

that the vector ’looks’ like the components appropriate to the rotated reference frame, and so we need to
find the projection of X onto the unit vectors that define the rotated frame. The details are shown in Fig.
(4b); notice that x2 = L1 + L2, L1 = x1 tanθ , and x02 = L2cosθ . From this it follows that
x02 = (x2 − x1tanθ )cosθ = −x1 sinθ + x2 cosθ . By a similar calculation we can find that
x01 = x1cos(θ ) + x2 sin(θ ). The component x03 that is aligned with the axis of the rotation vector is
unchanged, x03 = x3 , and so the set of equations for the primed components may be written as a column
vector  0  
x1 cos θ + x2 sin θ

x1
X0 =  x02  =  −x1 sin θ + x2 cos θ  . (9)
0
x3 x3
By inspection this can be factored into the product
X0 = RX, (10)
where X is the matrix of stationary frame components and R is the rotation matrix,

cos θ sin θ 0

R(θ ) =  − sin θ cos θ 0  . (11)


0 0 1
2 NONINERTIAL REFERENCE FRAMES 16

This θ is the angle displaced by the rotated reference frame and is positive counterclockwise. The
position vector observed from the rotated frame will be denoted by X0 ; to construct X0 we sum the
rotated components, X0 , times a set of unit vectors that are fixed and thus

X0 = e1 x01 + e2x02 + e3 x03 = EX0 (12)

For example, the position vector X of Fig. (4) is at an angle of about 45◦ counterclockwise from
the e1 unit vector and the rotated frame is at θ = 30◦ counterclockwise from the stationary frame one.
That being so, the position vector viewed from the rotated reference frame, X0 , makes an angle of 45◦ -
30◦ = 15◦ with respect to the e1 (fixed) unit vector seen within the rotated frame, Fig. (5). As a kind of
verbal shorthand we might say that the position vector has been ’transformed’ into the rotated frame by
Eqs. (9) and (12). But since the vector has an objective existence, what we really mean is that the
components of the position vector are transformed by Eqn. (9) and then summed with fixed unit vectors
to yield what should be regarded as a new vector, X0 , the position vector that we observe from the
rotated (or rotating) reference frame.14
Velocity: The velocity of parcel P seen in the stationary frame is just the time rate of change of the
position vector seen in that frame,
dX d dX
= EX = E ,
dt dt dt
since E is time-independent. The velocity of parcel P as seen from the rotating reference frame is
similarly
dX0 d dX0
= EX0 = E ,
dt dt dt
which indicates that the time derivatives of the rotated components are going to be very important in
what follows. For the first derivative we find
dX0 d(RX) dR dX
= = X+R . (13)
dt dt dt dt
14 If
the somewhat formal-looking Eqs. (9) through (12) do not have an immediate and concrete meaning for you, then
the remainder of this important section will probably be a loss. Some questions/assignments to help you along: 1) Verify
Eqs. (9) and (12) by some direct experimentation, i.e., try them and see. 2) Show that the transformation of the vector
components given by Eqs. (10) and (11) leaves the magnitude of the vector unchanged, i.e., |X0 | = |X|. 3) Verify that
R(θ1 )R(θ2 ) = R(θ1 + θ2 ) and that R(θ )−1 = R(−θ ), where R−1 is the inverse (and also the transpose) of the rotation
matrix. 4) Show that the unit vectors that define the rotated frame can be related to the unit vectors of the stationary frame by
`E = ER−1 and hence the unit vectors observed from the stationary frame turn the opposite direction of the position vector
observed from the rotating frame (and thus the reversed prime). The components of an ordinary vector (a position vector
or velocity vector) are thus said to be contravariant, meaning that they rotate in a sense that is opposite the rotation of the
−1
coordinate system. What, then, can you make of `EX0 = ER RX? A concise and clear reference on matrix representations
of coordinate transformations is by J. Pettofrezzo Matrices and Transformations (Dover Pub., New York, 1966). An excellent
all-around reference for undergraduate-level applied mathematics including coordinate transformations is by M. L. Boas,
Mathematical Methods in the Physical Sciences, 2nd edition (John Wiley and Sons, 1983).
2 NONINERTIAL REFERENCE FRAMES 17

1.2 1.2

1 1 e
e 2
2
‘e2
0.8 0.8

P
0.6
0.6
X ‘e1
0.4
0.4

ψ P
0.2
X’
0.2

θ e
0
1 ψ e
a 0 b
1

−0.2
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2 −0.2
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2
stationary ref frame rotated ref frame

Figure 5: (a) The position vector X seen from the stationary reference frame. (b) The position vector as
seen from the rotated frame, denoted by X0 . Note that in the rotated reference frame the unit vectors are
labeled ei since they are fixed; when these unit vectors are seen from the stationary frame, as on the left,
they are labeled `ei . If the position vector is stationary in the stationary frame, then θ + ψ = constant.
The angle ψ then changes as d ψ /dt = −d θ /dt = −Ω, and thus the vector X0 appears to rotate at the
same rate but in the opposite sense as does the rotating reference frame.

The second term on the right side of Eqn. (13) represents velocity components from the stationary
frame that have been transformed into the rotating frame, as in Eqn. (10). If the rotation angle θ was
constant so that R was independent of time, then the first term on the right side would vanish and the
velocity components would transform exactly as do the components of the position vector. In that case
there would be no Coriolis force.
When the rotation angle is time-varying, as it will be here, the first term on the right side of Eqn.
(13) is non-zero and represents a velocity component that is induced solely by the rotation of the
reference frame. For an Earth-attached reference frame
θ = θ0 + Ωt,
where Ω is Earth’s rotation rate measured with respect to the distant stars, effectively a constant defined
below (and θ0 is unimportant). Though Ω may be presumed constant, the associated reference frame is
nevertheless accelerating and is noninertial in the same way that circular motion at a steady speed is
accelerating because the direction of the velocity vector is continually changing. Given this θ (t), the
time-derivative of the rotation matrix is

− sin θ (t) cos θ (t) 0

dR
= Ω  − cos θ (t) − sin θ (t) 0  , (14)
dt
0 0 0
2 NONINERTIAL REFERENCE FRAMES 18

which, notice, this has all the elements of R, but shuffled around. By inspection, this matrix can be
factored into the product of a matrix C and R as
dR
= Ω CR(θ (t)), (15)
dt
where the matrix C is    
0 1 0 1 0 0
C =  −1 0 0  =  0 1 0  R(π /2). (16)
0 0 0 0 0 0
Multiplication by C acts to knock out the component ( )3 that is parallel to Ω and causes a rotation of
π /2 in the plane perpendicular to Ω . Substitution into Eqn. (13) gives the velocity components
appropriate to the rotating frame
d(RX) dX
= ΩCRX + R , (17)
dt dt
or using the prime notation ( )0 to indicate multiplication by R, then
 0
dX0 0 dX
= ΩCX + (18)
dt dt
The second term on the right side of Eqn. (18) is just the rotated velocity components and is present
even if Ω vanished (a rotated but not a rotating reference frame). The first term on the right side
represents a velocity that is induced by the rotation rate of the rotating frame; this induced velocity is
proportional to Ω and makes an angle of π /2 radians to the right of the position vector in the rotating
frame (assuming that Ω > 0).
To calculate the vector form of this term we can assume that the parcel P is stationary in the
stationary reference frame so that dX/dt = 0. Viewed from the rotating frame, the parcel will appear to
move clockwise at a rate that can be calculated from the geometry (Fig. 6). Let the rotation in a time
interval δ t be given by δ ψ = −Ωδ t; in that time interval the tip of the vector will move a distance
|δ X0 | = |X0 |sin(δ ψ ) ≈ | X0 |δ ψ , assuming the small angle approximation for sin(δ ψ ). The parcel will
move in a direction that is perpendicular (instantaneously) to X0 . The velocity of parcel P as seen from
0
the rotating frame and due solely to the coordinate system rotation is thus limδ t→0 δδXt = − ×X0 , the
vector cross-product equivalent of ΩCX0 (Fig. 7). The vector equivalent of Eqn. (18) is then
 0
dX0 0 dX
= − ×X + (19)
dt dt

The relation between time derivatives given by Eqn. (19) is general; it applies to all vectors, e.g.,
velocity vectors, and moreover, it applies for vectors defined at all points in space. Hence the
relationship between the time derivatives may be written as an operator equation,

d( )0 d( ) 0
 
0
= − Ω×
Ω×( ) + (20)
dt dt
2 NONINERTIAL REFERENCE FRAMES 19

1 e2 Figure 6: The position vector X0 seen from


the rotating reference frame. The unit vectors
0.8
that define this frame, `ei , appear to be station-
δX
ary when viewed from within this frame, and
0.6
hence we label them with ei (not primed). As-
X’(t)
0.4
sume that Ω > 0 so that the rotating frame is
turning counterclockwise with respect to the
δψ
0.2
X’(t+δt) stationary frame, and assume that the parcel P
e1 is stationary in the stationary reference frame
0 so that dX/dt = 0. Parcel P as viewed from
the rotating frame will then appear to move
−0.2
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
clockwise on a circular path.
rotating ref. frame

that is valid for all vectors.15 From left to right the terms are: 1) the time rate of change of a vector as
seen in the rotating reference frame, 2) the cross-product of the rotation vector with the vector and 3)
the time rate change of the vector as seen in the stationary frame and then rotated into the rotating frame.
One way to describe Eqn. (20) is that the time rate of change and prime operators do not commute, the
difference being the cross-product term which, notice, represents a time rate change in the direction of
the vector, but not its magnitude. Term 1) is the time rate of change that we observe directly or seek to
solve when working from the rotating frame.
Acceleration: Our goal here is to relate the accelerations seen in the two reference frames and so
differentiating Eqn. (18) once more and after rearrangement of the kind used above
 2 0
d 2 X0 dX0 2 2 0 d X
2
= 2ΩC +Ω C X + (21)
dt dt dt 2

The middle term on the right includes multiplication by the matrix C2 = CC,
       
1 0 0 1 0 0 1 0 0 1 0 0
C2 =  0 1 0  R(π /2)  0 1 0  R(π /2) =  0 1 0  R(π ) = −  0 1 0  ,
0 0 0 0 0 0 0 0 0 0 0 0
that knocks out the component corresponding to the rotation vector Ω and reverses the other two
components; the vector equivalent of Ω2 C2 X0 is thus −Ω Ω ×X0 (Fig. 7). The vector equivalent of
Ω ×Ω
15 Imagine arrows taped to a turntable with random orientations. Once the turntable is set into (solid body) rotation, all of
the arrows will necessarily rotate at the same rotation rate regardless of their position or orientation. The rotation will, of
course, cause a translation of the arrows that depends upon their location, but the rotation rate is necessarily uniform, and
this holds regardless of the physical quantity that the vector represents. This is of some importance for our application to a
rotating Earth, since Earth’s motion includes a rotation about the polar axis, as well as an orbital motion around the Sun and
yet we represent Earth’s rotation by a single vector.
2 NONINERTIAL REFERENCE FRAMES 20

Ω Figure 7: A schematic showing the relation-


ship of a vector X, and various cross-products
with a second vector Ω (note the signs). The
vector X is shown with its tail perched on the
axis of the vector Ω as if it were a position
−Ω x Ω x X vector. This helps to visualize the direction of
the cross-products, but it is important to note
that the relationship among the vectors and
vector products shown here holds for all vec-
−Ω x X tors, regardless of where they are defined in
X space or the physical quantity, e.g., position
or velocity, that they represent.

Eqn. (21) is then16


 2 0
d 2 X0 dX0 0 d X
= − Ω
2Ω× − Ω Ω
×Ω ×X + (22)
dt 2 dt dt 2
Note the similarity with Eqn. (3). From left to right the terms are 1) the acceleration as seen in the
rotating frame, 2) the Coriolis term, 3) the centrifugal17 term, and 4) the acceleration as seen in the
stationary frame and then rotated into the rotating frame. As before, term 1) is the acceleration that we
directly observe or seek to solve for when working from the rotating reference frame.

2.2.2 Stationary ⇒ Inertial; Rotating ⇒ Earth-Attached

The third and final step in this derivation of the Coriolis force is to define the inertial reference frame
one, and then the rotation rate of frame two. To make frame one inertial it is presumed that the unit
vectors ei could in principle be aligned on the distant, ’fixed stars’.18 The rotating frame two is
16 The relationship between the stationary and rotating frame velocity vectors given by Eqs. (18) and (19) is clear visually
and becomes intuitive given just a little experience. It is not so easy to intuit the corresponding relationship between the
accelerations given by Eqs. (21) and (22). To understand the transformation of acceleration there is really no choice but to
understand (be able to reproduce and then explain) the mathematical steps going from Eqn. (18) to Eqn. (21) and/or from
Eqn. (19) to Eqn. (22).
17 ’Centrifugal’ and ’centripetal’ have Latin roots, centri+fugere and centri+peter, meaning center-fleeing and center-

seeking, respectively. Taken literally they would indicate the sign of a radial force, for example. However, they are very
often used to mean the specific term ω 2 r, i.e., centrifugal force when it is on the right side of an equation of motion and
centripetal acceleration when it is on the left side.
18 This is aside from celestial rotation. Even so ‘fixed’ is a matter of degree; certainly the Sun and the planets do not qualify,

but even some nearby stars move detectably over the course of a year. The intent is that the most distant stars should serve as
2 NONINERTIAL REFERENCE FRAMES 21

presumed to be attached to Earth, and the rotation rate Ω is then given by the rate at which the same
fixed stars are observed to rotate overhead, one revolution per sidereal day (Latin for from the stars), 23
hrs, 56 min and 4.09 sec, or
Ω = 7.2921×10−5 rad s−1 . (23)
A sidereal day is only about four minutes less than a solar day, and so in a purely numerical sense,
Ω ≈ Ωsolar = 2π /24 hours = 7.2722×10−5 rad s−1 which is certainly easier to remember than is Eqn.
(23). For the purpose of a rough estimate, the small numerical difference between Ω and Ωsolar is not
significant. However, the difference between Ω and Ωsolar can be told in numerical simulations and in
well-resolved field observations. On Mach’s Principle,18 the difference between Ω and Ωsolar is, of
course, highly significant.
Earth’s rotation rate is very nearly constant, and the axis of rotation maintains a nearly steady
bearing on a point on the celestial sphere that is close to the North Star, Polaris (cover graphic). The
Earth’s rotation vector thus provides a definite orientation of Earth within the universe, and the rotation
rate has an absolute magnitude. The rotation rate sensors noted in footnote 10 read out Earth’s rotation
rate with respect to the fixed stars as a kind of gage pressure, called ’Earth rate’.19

sign posts for the spatially-averaged mass of the universe as a whole on the hypothesis that inertia arises whenever there is an
acceleration (linear or rotational) with respect to the mass of the universe. This grand idea was expressed most forcefully by
the Austrian philosopher and physicist Ernst Mach, and is often termed Mach’s Principle (see, e.g., J. Schwinger, Einstein’s
Legacy Dover Publications, 1986; M. Born, Einstein’s Theory of Relativity, Dover Publications, 1962). Mach’s Principle
seems to be in accord with all empirical data, including the magnitude of the Coriolis force. Mach’s principle is best thought
of as a relationship, and is not, in and of itself, the fundamental mechanism of inertia. A new hypothesis takes the form of so-
called vacuum stuff (or Higgs field) that is presumed to pervade all of space and so provide a local mechanism for resistance
to accelerated motion (see P. Davies, ‘On the meaning of Mach’s principle’, http://www.padrak.com/ine/INERTIA.html). The
debate between Newton and Leibniz over the reality of absolute space — which had seemed to go in favor of relative space,
Leibniz and Mach’s Principle — has been renewed in the search for a physical origin of inertia. If this can be achieved, we
can then point to a physical origin of the Coriolis force.
Observations on the fixed stars are a very precise means to define rotation rate, but can not, in general, be used to define the
linear translation or acceleration of a reference frame. The only way to know if a reference frame that is aligned on the fixed
stars is inertial is to carry out mechanics experiments and test whether Eqn.(1) holds and global momentum is conserved. If
yes, the frame is inertial.
19 For the present purpose Ω may be presumed constant. In fact, there are small but observable variations of Earth’s rotation

rate due mainly to changes in the atmospheric and oceanic circulation and due to mass distribution within the cryosphere,
see B. F. Chao and C. M. Cox, ‘Detection of a large-scale mass redistribution in the terrestrial system since 1998,’ Science,
297, 831–833 (2002), and R. M. Ponte and D. Stammer, ‘Role of ocean currents and bottom pressure variability on seasonal
polar motion,’ J. Geophys. Res., 104, 23393–23409 (1999). The direction of Ω with respect to the celestial sphere also varies
detectably on time scales of tens of centuries on account of precession, so that Polaris has not always been the pole star (Fig.
), even during historical times. The slow variation of Earth’s orbital parameters (slow enough to be assumed to vanish for
our purpose) are an important element of climate, see e.g., J. A. Rial, ‘Pacemaking the ice ages by frequency modulation of
Earth’s orbital eccentricity,’ Science, 285, 564–568 (1999).
As well, Earth’s motion within the solar system and galaxy is much more complex than a simple spin around a perfectly
stable polar axis. Among other things, the Earth orbits the Sun in a counterclockwise direction with a rotation rate of 1.9910
×10−7 s−1 , which is about 0.3% of the rotation rate Ω. Does this orbital motion enter into the Coriolis force, or otherwise
2 NONINERTIAL REFERENCE FRAMES 22

Assume that the inertial frame equation of motion is

d 2X d 2X
M = F + G ∗ M and M = F + g∗ M (24)
dt 2 dt 2
(G∗ is the component matrix of g∗). The acceleration and force can always be viewed from another
reference frame that is rotated (but not rotating) with respect to the first frame,
 2 0  2 0
d X 0 0 d X
2
M = F + G∗M and M = F0 + g0∗ M, (25)
dt dt 2
as if we had chosen a different set of fixed stars or multiplied both sides of Eqn. (22) by the same
rotation matrix. This equation of motion preserves the global conservation and Galilean transformation
properties of Eqn. (24). To find the rotating frame equation of motion, eliminate the rotated acceleration
from Eqn. (25) using Eqs. (21) and (22) and then solve for the acceleration seen in the rotating frame:
the components are
d 2 X0 dX0
2
M = 2ΩC M − Ω2 C2X0 M + F0 + G0∗ M (26)
dt dt
and the vector equivalent is

d 2 X0 dX0
M = − Ω
2Ω × Ω ×X0 M + F0 + g∗0 M
M − Ω ×Ω (27)
dt 2 dt
Eqn. (27) has the form of Eqn. (4), the difference being that the noninertial reference frame is rotating
rather than translating. If the origin of this noninertial reference frame was also accelerating, then there
would be a third inertial force term, −(d 2Xo /dt 2)M. Notice that we are not yet at Eqn. (2); in Section
4.1 the centrifugal force and gravitational mass attraction terms will be are combined into the
time-independent inertial force g.

2.2.3 Remarks on the transformed equation of motion

Once the transformation rule for accelerations, Eqn. (22), is in hand, the path to the rotating frame
equation of motion is short and direct — if Eqn. (25) holds in a given reference frame (say an inertial
affect the dynamics of the atmosphere and oceans? The short answer is no and yes. We have already accounted for the rotation
of the Earth with respect to the fixed stars. Whether this rotation is due to a spin about an axis centered on the Earth or due to
a solid body rotation about a displaced center is not relevant for the Coriolis force per se, as noted in the discussion of Eqn.
(20). However, since Earth’s polar axis is tilted significantly from normal to the plane of the Earth’s orbit around the Sun (the
tilt implied by Fig. ()), we can ascribe Earth’s rotation Ω to spin alone. The orbital motion about the Sun combined with
Earth’s finite size gives rise to tidal forces, which are small but important spatial variations of the centrifugal/gravitational
balance that holds for the Earth-Sun and for the Earth-Moon as a whole (described particularly well by French8 ). A question
for you: What is the rotation rate of the Moon? Hint, make a sketch of the Earth-Moon orbital system and consider what we
observe of the Moon from Earth. What would the Coriolis and centrifugal forces be on the Moon?
2 NONINERTIAL REFERENCE FRAMES 23

frame, but that is not essential) then Eqs. (26) and (27) hold exactly in a frame that rotates at the
constant rate and direction given by Ω with respect to the first frame. The rotating frame equation of
motion includes two terms that are dependent upon the rotation vector, the Coriolis term,
−2ΩΩ×(dX0 /dt), and the centrifugal term, −Ω Ω ×ΩΩ ×X0 . These terms are sometimes written on the left
side of an equation of motion as if they were going to be regarded as part of the acceleration, i.e.,

d 2 X0 dX0
M + Ω
2Ω × Ω ×X0 M = F0 + g∗0 M.
M + Ω ×Ω (28)
dt 2 dt
Comparing the left side of Eqn. (28) with Eqn. (22), it is evident that the rotated acceleration is equal to
the rotated force,  2 0
d X
M = F0 + g∗0 M,
dt 2
which is well and true and the same as Eqn. (25).20 However, it is crucial to understand that the left side
of Eqn. (28), (d 2X/dt 2 )0 is not the acceleration that is observed from the rotating reference frame,
d 2X0 /dt 2 . When Eqn. (28) is solved for d 2 X0 /dt 2, it follows that the Coriolis and centrifugal terms are,
figuratively or literally, sent to the right side of the equation of motion where they are interpreted as if
they were forces.
When the Coriolis and centrifugal terms are regarded as forces — and it is argued here that they
should be when observing from a rotating reference frame — they have all of the peculiar properties of
inertial forces noted in Section 2.1. From Eqn. (28) (and Eqn. 4) it is evident that the centrifugal and
Coriolis terms are exactly proportional to the mass of the parcel observed, whatever that mass may be.
The acceleration associated with these inertial forces arises from the rotational acceleration of the
reference frame, combined with relative velocity for the Coriolis force. They differ from central forces
F and g∗M in the respect that there is no physical interaction that causes the Coriolis or centrifugal
force and hence there is no action-reaction force pair. As a consequence the rotating frame equation of
motion does not retain the global conservation of momentum that is a fundamental property of the
inertial frame equation of motion and central forces (an example of this nonconservation is described in
Section 3.4). Similarly, we note here only that invariance to Galilean transformation is lost since the
Coriolis force involves the velocity rather than velocity derivatives. Thus V0 is an absolute velocity in
the rotating reference frame of the Earth. If we need to call attention to these special properties of the
Coriolis force, then the usage Coriolis inertial force seems appropriate because it is free from the taint
of unreality that goes with ’virtual force’, ’fictitious correction force’, etc., and because it gives at least a
hint at the origin of the Coriolis force. It is important to be aware of these properties of the rotating
frame equation of motion, and also to be assured that in most analysis of geophysical flows they are of
no great practical consequence. What is most important is that the rotating frame equation of motion
20 Recallthat Ω = Ω0 and so we could put a prime on every vector in this equation. That being so, it would be better to
remove the visually distracting primes and simply note that the resulting equation holds in a steadily rotating reference frame.
We will keep the primes for now, since we will be considering both inertial and rotating reference frames until Section 5.
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 24

offers a very significant gain in simplicity compared to the inertial frame equation of motion, discussed
further in Section 4.
The Coriolis and centrifugal forces taken individually have simple interpretations. From Eqn. (27)
it is evident that the Coriolis force is normal to the velocity, dX0 /dt, and to the rotation vector, Ω . The
Coriolis force will thus tend to cause the velocity to change direction but not magnitude, and is
appropriately termed a deflecting force as noted in Section 1. The centrifugal force is in a direction
perpendicular to and directed away from the axis of rotation. Notice that the Coriolis force is
independent of position, while the centrifugal force clearly is not. The centrifugal force is independent
of time. How these forces effect dynamics in simplified conditions will be considered further in
Sections 3, 4.3 and 5.

3 Inertial and noninertial descriptions of elementary motions

The equations of motion (24) and (27) will be evaluated for several prescribed, elementary motions
whose inertial frame dynamics will be very familiar. The object will be to compute the description of
these motions as they would be observed from a noninertial reference frame, and the goal will be to
understand how the accelerations and the inertial forces — gravity, centrifugal and Coriolis — depend
upon the reference frame. Though the motions considered here are truly elementary, nevertheless the
analysis is slightly subtle in that the inertial force terms will change identity, as if be fiat, when changing
the reference frame. To appreciate that there is more to this than an arbitrary relabeling of terms it will
be very helpful to make a sketch of each case, starting with the observed motion (acceleration).
There is an important distinction to be made between what will be termed contact forces, F, that
act over the surface of the parcel, and the acceleration due to gravity, gM, which is an inertial force that
acts throughout the body of the parcel (in this section there is no distinction between g and g∗) (Table 1).
To measure the contact forces, the parcel could be enclosed in a wrap-around strain gage that reads out
the vector sum of the tangential and normal stresses acting on the surface of the parcel. To measure the
gravity field, one could observe the direction of a stationary plumb line, and the magnitude inferred
from its period of oscillation.12

3.1 Switching sides

In this section the equations of motion, Eqn. (24) and (27), are evaluated for several truly elementary
motions. The analysis is slightly subtle nevertheless insofar as the terms that represent accelerations and
inertial forces will seem to change identity, as if by fiat, when we change reference frames. To
understand that there is more to this than simply relabeling and reinterpreting terms in an arbitrary way,
it will be very helpful for you to make a sketch of each case described below, beginning with the
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 25

A characterization of the forces on geophysical flows.

central? inertial? Galilean invariant? position only?


contact forces yes no yes no
grav. mass attraction yes yes yes yes
centrifugal no yes yes yes
Coriolis no yes no no

Table 1: Contact forces on fluid parcels include pressure gradients (normal to a surface) and frictional
forces (mainly tangential to a surface). The centrifugal force noted here is that associated with Earth’s ro-
tation. ’position only’ means dependent upon the parcel position but not the parcel velocity, for example.
This table ignores electromagnetic forces that are usually small.

acceleration.
Consider a parcel of known, fixed mass M that is at rest and in contact with the ground, say, in a
reference frame where the acceleration of gravity is known from independent observations, e.g.,
pendulum experiments. The strain gauge will read out a contact force Fz , which is upwards, from the
perspective of the parcel. The vertical component of the equation of motion for the parcel is

d 2z
M = Fz − g,
dt 2
where the observable acceleration is written on the left side of the equation of motion (even when
regarded as known) and the forces are listed on the right side. In this case the acceleration vanishes and
so
0 = Fz − gM, (29)
which indicates a static force balance between the upward contact force, Fz , and the downward force
due to gravity. Now suppose that the same parcel is observed from a reference frame that is in free-fall
and accelerating downwards at the rate −g.21 When viewed from this reference frame, the parcel is
accelerating upward at the rate g that is just the complement of the acceleration of the free-falling frame.
In this frame there is no gravitational force (imagine astronauts floating in space and attempting
pendulum experiments) and so the only force recognized as acting on the parcel is the contact force, Fz ,
21 Gravitational mass attraction is an inertial force and a central force that has a very long range. Consider two gravitating
bodies and a reference frame attached to one of them, say parcel one, which will then be observed to be at rest. If parcel two
is then found to accelerate towards parcel one, the total momentum of the system (parcel one plus parcel two) will not be
conserved, i.e., in effect, gravity would not be recognized as a central force. A reference frame attached to one of the parcels
is thus noninertial. To define an inertial reference frame in the presence of mutually gravitating bodies we can use the center
of mass of the system, and then align on the fixed stars. This amounts to putting the entire system into free-fall with respect
to any larger scale (external to this system) gravitational mass attraction (for more on gravity and inertial reference frames see
http://plato.stanford.edu/entries/spacetime-iframes/).
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 26

which is unchanged from the case before. The equation of motion for the parcel observed from this
free-falling reference frame is then
d 2z0
M = Fz ,
dt 2
or evaluating the acceleration, d 2z0 /dt 2 = g,

gM = Fz . (30)

Notice that in going from Eqn. (29) to the free-falling frame Eqn. (30) the contact force Fz is unchanged
(invariant) while the term involving g has switched sides; gM is an inertial force in the reference frame
appropriate to Eqn. (29) and is transformed into an acceleration in the free-falling reference frame
described by Eqn. (30). The equation of motion makes perfectly good sense either way, but what we
observe as an acceleration in one frame can appear to be an inertial force in another reference frame.
Exactly the same kind of switching sides happens also with centrifugal and Coriolis inertial forces.
Now consider just the horizontal motion of this parcel, so that gravity and the vertical component
of the motion will be ignored. Presume that F = 0 (no horizontal contact force) and hence the inertial
frame equation of motion expanded in polar coordinates (derived in Section 3.4 and repeated here for
convenience),

d 2X
 2   
d r 2 dr dω
M = − rω Mer + 2ω + r Meθ = Fr er + Fθ eθ ,
dt 2 dt 2 dt dt

vanishes term by term when the parcel is at rest in the inertial frame. Suppose that the same parcel is
viewed from a steadily rotating reference frame and that it is at a distance r0 from the origin of the
rotating frame. Viewed from this frame, the parcel will appear to be moving in a circle of radius
r0 = constant and in a direction opposite the rotation of the reference frame, ω 0 = −Ω, just as in Figure
(6). The velocity is V0 = −ΩΩ ×X0 . The rotating frame equation of motion in polar coordinates is then

d 2 X0
 2 0   0 0

d r 0 02 0 0 dr 0 dω
2
M = 2
−r ω Mer + 2ω +r Me0θ
dt dt dt dt
dr0
 
= r Ω M + 2Ωω r M + Fr er + −2Ω M + Fθ e0θ .
0 2 0 0 0 0 0

(31)
dt

The strain gage can be read from this rotating frame and it is still the case that Fr0 = Fθ0 = 0. All of the
azimuthal component terms vanish individually, but three of the radial component terms are nonzero,

−r0 ω 02 = r0 Ω2 + 2Ωω 0 r0 , (32)

and indicate an interesting balance between the centripetal acceleration, −r0 ω 02 (the observed
acceleration is listed on the left hand side), and the sum of the centrifugal and Coriolis inertial forces
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 27

(the right hand side, divided by M, and note that ω 0 = −Ω).22 Interesting perhaps, but disturbing as
well; a parcel that was at rest in an inertial frame has acquired a rather complex momentum balance
when observed from a rotating reference frame. It is sorely tempting to deem the Coriolis and
centrifugal terms to be ’virtual’, or ’fictitious, correction’ forces to acknowledge this discomfort.5 To be
consistent, we would have to do the same for the observed, centripetal acceleration on the left hand side.
Labeling terms this way wouldn’t add anything useful, and it might serve to obscure the fundamental
issue — that all accelerations and inertial forces are relative to a reference frame. As found in the
example of a free-falling reference frame, this applies equally to gravitational mass attraction and to
centrifugal and Coriolis forces.

3.2 To get a feel for the Coriolis force

The centrifugal force is something that we encounter often in daily life. For example, a runner having
V = 5 m s−1 and making a moderately sharp turn, radius R = 15 m, will easily feel the centrifugal force,
(V 2/R)M ≈ 0.15gM, and will compensate instinctively by leaning toward the center of the turn. It is
unlikely that a runner would think of this centrifugal force as virtual or fictitious.
The Coriolis force associated with Earth’s rotation is by comparison very small, only about
2ΩV M ≈ 10−4gM for the same runner. To experience the Coriolis force in the same direct way that we
can feel the centrifugal force, i.e., to feel it in our bones, will thus require a platform having a rotation
rate that exceeds Earth’s rotation rate by a factor of about 104. A typical merry-go-round having a
rotation rate Ω = 2π /12 rad s−1 = 0.5 rad s−1 is ideal. To calculate the forces we will need to a body
mass, say M = 75 kg, the standard airline passenger before the era of super-sized meals and passengers.

3.2.1 Zero relative velocity

To start, let’s presume that we are sitting quietly near the outside radius r = 6 m of a merry-go-round
that it is rotating at a steady rate, Ω = 0.5 rad s−1 . How does the momentum balance of our motion
depend upon the reference frame?
Viewed from an approximate inertial frame outside of the merry-go-round (fixed stars are not
required given the rapid rotation rate), the polar coordinate momentum balance Eqn. (31) with ω = Ω
and dr/dt = d ω /dt = Fθ = 0 reduces to a two term radial balance,

−rΩ2 M = Fr , (33)
22 Two problems for you: 1) Given the polar coordinate velocity, Eqn. (42), show that Eqn. (31) can be derived also from
the vector form of the equation of motion, Eqn. (27). 2) Sketch the balance of forces in a case where the rotation rate Ω is
positive and then again where it is negative. Is this consistent with Eqn. (32)?
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 28

in which a centripetal acceleration (×M) is balanced by an inward-directed radial (contact) force, Fr . We


can readily evaluate the former and find −rΩ2 M = Fr = −112 N, which is equal to the weight on a
mass of Fr /g = 11.5 kg for a nominal g. This is just what the strain gauge (the one on the seat of your
pants) reads out.
Viewed from the rotating reference frame, i.e., our seat on the merry-go-round, we are stationary
and of course not accelerating. To evaluate the rotating frame momentum equation, Eqn. 31, set
ω 0 = 0, r0 = constant leaving a two term radial force balance,

0 = r0 Ω2 M + Fr0 . (34)

The physical conditions are unchanged and thus the strain gage reads out exactly as before, and
Fr0 = Fr = −112 N. What has changed is that the term r0 Ω2 M, an acceleration in the inertial frame, is
now on the right side of the momentum equation and is regarded as the centrifugal force. Within the
rotating frame, the centrifugal force is quite vivid; it appears that we are being pushed outwards, or
centrifugally, by an inertial force that is opposed by the centripetal contact force Fr0 . This is exactly the
relationship between weight and a contact force described in Section 3.1. The centrifugal force
produces a radial acceleration on every stationary object that depends only upon the radius, r0 . For
example, a plumb line makes an angle to the vertical of arctan(r0 Ω2 /g), where the vertical direction
and g are in the absence of rotation. A centrifugal force thus contributes to the direction and magnitude
of the time-independent acceleration field observed in a rotating frame, an important point returned to in
Section 4.1.

3.2.2 With relative velocity

Most merry-go-rounds have signs posted which caution riders to remain in their seats once the ride
begins. This is a good and prudent rule, of course. But if the goal is to get a feel for the Coriolis force
then we may decide to go for a (very cautious) walk on the merry-go-round. Azimuthal relative
velocity: Let’s assume that we walk azimuthally so that r = 6 m and constant. A reasonable walking
pace under the circumstance is about Uw = 1.5 m s−1 , which corresponds to a relative rotation rate
ω 0 = 0.25 rad s−1 , and recall that Ω = 0.5 rad s−1 . If the direction is in the direction of the
merry-go-round rotation, then ω = Ω + ω 0 = 0.75 rad s−1 . From the inertial frame momentum
equation (31), the centripetal force required to maintain r = constant at this greater angular velocity is

−rω 2 M = −r(Ω + ω 0 )2 M = Fr ≈ −253 N,

or roughly twice the force required when seated. If we then reverse direction and walk at the same
speed against the rotation of the merry-go-round, Fr is reduced to about -28 N. This pronounced
variation of Fr with ω 0 is a straightforward consequence of the quadratic dependence of centripetal
acceleration upon the rotation rate, ω .
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 29

When this motion is viewed from the rotating frame of the merry-go-round, we distinguish
between the rotation rate of the merry-go-round, Ω, and the relative rotation rate, ω 0 , due to our walking
speed. The radial component of the rotating frame momentum equation reduces to

−r0 ω 02 M = (r0 Ω2 + 2r0 Ωω 0 )M + Fr0 . (35)

The term on the left is the comparatively small centripetal acceleration; the first term on the right is the
usual centrifugal force, and the second term on the right, 2r0 Ωω 0 , is the Coriolis force. The Coriolis
force is substantial, 2r0 Ωω 0 M ± 112 N, with the sign determined by the direction of our motion relative
to Ω. If Ω > 0 and ω 0 > 0 then the Coriolis force is positive and radial and to the right of and normal to
the azimuthal relative velocity. Given what we found in the previous paragraph, it is tempting to
identify the Coriolis force as the (relative)velocity-dependent part of the centrifugal force. This is,
however, somewhat loose and approximate; loose because the centrifugal force is defined to be
dependent upon rotation rate and position only and approximate because this ignores the small
centripetal acceleration term.
Radial relative velocity: Now consider a (very cautious) walk in the radial direction. To isolate the
effects of radial motion we will presume that our radial speed is constant at dr0 /dt = 1 m s−1 and that
we walk along a radial line so that our rotation rate also remains constant at ω = Ω. (In practice this is
difficult to maintain for more than a few steps, but that will suffice.) The resulting contact force F is
then in the azimuthal direction, and its magnitude and sense can most easily be interpreted in terms of
the balance of angular momentum, A = ω r02 M. In this circumstance the rate of change of angular
momentum A has been fully specified,
dA dr0
= 2Ωr0 M = r0 Fθ ,
dt dt
and must be accompanied by an azimuthal torque, r0 Fθ , exerted by the merry-go-round via an azimuthal
contact force.
Viewed from an inertial frame, the azimuthal component of the momentum balance, Eqn. (31),
reduces to
dr
2Ω M = Fθ , (36)
dt
where Fθ ≈ −75 N for the given data. This azimuthal contact force Fθ has the form of the Coriolis force,
but remember that we are viewing the motion from an inertial frame so that there is no Coriolis force. If
the radial motion was inward so that dr/dt < 0, then Fθ must be negative, or opposite the direction of
the merry-go-round rotation, since our angular momentum is necessarily becoming less positive. (Be
sure that these signs are clear before going on to consider this motion from the rotating frame.)
From within the rotating frame, the momentum equation reduces to an azimuthal force balance
dr0
0 = −2Ω M + Fθ0 , (37)
dt
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 30

0
where −2Ω dr 0
dt M is the Coriolis force and Fθ = Fθ as before. The contact force exerted by the
merry-go-round, Fθ0 , is balanced by an inertial force, the Coriolis force, in the direction opposed to Fθ0 .
0 dr 0
For example, if our radial motion is inward, dr dt ≤ 0, then the Coriolis force, −2Ω dt M ≥ 0, is to the
right of and normal to our relative velocity, just as we would have expected from the vectorial Coriolis
force. Notice that this interpretation of a Coriolis force is exactly the interpretation of centrifugal force
in the example of steady, circular motion and Eqn. (34): an acceleration seen from an inertial frame
appears to be an inertial force when viewed from the rotating frame.
Be careful! If you have a chance to do this experiment you will learn with the first step whether the
Coriolis force is better described as a real force or as a fictitious correction force. Be sure to ask
permission of the operator before you start walking around, and exercise genuine caution. The Coriolis
force is an inertial force and so is distributed throughout your body, unlike the contact force which acts
only where you are in contact with the merry-go-round, i.e., through the soles of your sneakers or by a
secure hand grip. The radial Coriolis force associated with azimuthal motion is much like an increase or
slackening of the centrifugal force and so is not particularly difficult to compensate. However, the
azimuthal Coriolis force associated with radial motion is quite surprising, even presuming that you are
the complete master of this analysis. If you do not have access to a merry-go-round or if you feel that
this experiment is unwise, then see Stommel and Moore9 for alternate ways to accomplish some of the
same things.

3.3 An elementary projectile problem

A very simple projectile problem can reveal some other aspects of rotating frame dynamics. Assume
that a projectile is launched at a speed U0 and at an angle to the ground β from a location [x y] = [0 y0 ].
The only force presumed to act on the projectile after launch is the downward force of gravity, −gMe3 ,
which is the same in either reference frame.

3.3.1 From the inertial frame

The equations of motion and initial conditions in Cartesian components are linear and uncoupled;

d 2x dx
2
= 0; x(0) = 0, = 0, (38)
dt dt

d 2y dy
2
= 0; y(0) = y0 , = U0 cos β ,
dt dt

d 2z dz
2
= −g; z(0) = 0, = U0 sin β ,
dt dt
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 31

1
inertial

0.5 0.5
inertial
rotating
z

y
0
1 0 rotating
0.5 1
Coriolis 0.5 Coriolis
0 0
y −0.5 −0.5 x −0.5
−0.5 0 0.5 1
x

Figure 8: Trajectory of a parcel launched with a horizontal velocity in the positive y-direction as seen
from an inertial reference frame (solid line, displaced in the y-direction only), and as seen from a rotating
frame (dashed, curves lines). The left and right panels are 3-dimensional and plan views. The dotted
curve is with the Coriolis force only (the motivation for this is in Section 4). This trajectory has the form
of a circle, and if the projectile had not returned to the surface, z = 0, it would have made a complete loop
back to the starting point. Some excellent videos from comparable laboratory experiments are found at
http://planets.ucla.edu/featured/spinlab-geoscience-educational-film-project/

2U0 sin β
where M has been divided out. The solution for 0 < t < g , the time the parcel is in the air, is

x(t) = 0, (39)
y(t) = y0 + tU0 cos β ,
1
z(t) = t(U0 sin β − gt)
2
is in Fig. (8).

3.3.2 From the rotating frame

How would this same motion look when viewed from a rotating reference frame? And, how could we
compute the motion from within a rotating reference frame?
The first question can be answered most directly by rotating the trajectory, Eqn. (39), via the
rotation matrix, Eqn. (12), X0 = RX with θ = Ωt, and with the result

x0 (t) = (y0 + tU0 cos β ) sin (Ωt), (40)

y0 (t) = (y0 + tU0 cos β ) cos (Ωt),


3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 32

1 Figure 9: (a) The distance from


a the origin in the horizontal plane
0.8 inertial and
dist from origin

rotating
0.6
for the trajectories of Fig. (8).
0.4
The distance from the origin is
Coriolis
identical for the inertial and rotat-
0.2
ing trajectories, and reduced for
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
the Coriolis trajectory (discussed
time in Section 4). (b) The distance
1.5 along the path in the horizontal
plane for the same trajectories.
b
The slope gives the speed of the
dist along path

1 rotating

parcel. The inertial and Coriolis


0.5 frame trajectories retain their ini-
inertial and
Coriolis
tial speed and are identical; the
0 rotating frame trajectory acceler-
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
time ates due to the centrifugal force.

1
z0 (t) = z = t(U0 sin β − gt),
2
valid over the time interval as before. The z component is unchanged in going to the rotating reference
frame since the rotation axis was aligned with z. This is quite general; motion that is parallel to the
rotation vector Ω is unchanged, just as we would expect from the three-dimensional vector Coriolis
force. On the other hand, motion in the (x, y)-plane perpendicular to the rotation vector can be altered
quite substantially, depending upon the phase Ωt. In the case shown in Fig. (8), the change of phase is
2.0 at the end of the trajectory, so that rotation effects are prominent.23 One important aspect of the
trajectory is not changed, however, the (horizontal) radius,
p p
x02 + y02 = x2 + y2 ,

since the coordinate systems have coincident origins (Fig. 9a)


How could we compute the trajectory in the rotating frame? The Cartesian component equations in
the rotating frame are a bit awkward (the x-component only):

d 2 x0 dy0 2 x02
= 2Ω + Ω .
dt 2
p
dt x02 + y02
An elementary problem in the inertial frame transforms into a pair of coupled, nonlinear equations in
the rotating frame (z0 = z). We can always solve these equations numerically, but we are better off in
23 A well-thrown baseball travels at about 45 m s−1 . How much will it be deflected as it travels over a distance of 30 m?
Use Earth’s rotation rate given by Eqn. (23)(as we will see in Section 4.2 this is appropriate for the north pole). A long-range
artillery shell has an initial speed of about 700 m s−1 . Assuming the shell is launched at angle to the ground of 30 degrees,
how much will it be deflected over its trajectory (ignoring air resistance)?
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 33

this and many problems involving rotation to use cylindrical polar coordinates where we can take
advantage of what we have already learned about the rotating frame solution. We know that

r0 = r = y0 + tU0 cos β ,

and that the angle in the inertial frame, θ , is constant in time since the motion is purely radial and for
the specific case considered, θ = π /2. The rotation rates are related by ω 0 = −Ω, and thus

θ 0 = π /2 − Ωt.

Both the radius and the angle increase linearly in time, and the horizontal trace of the trajectory as seen
from the rotating frame is Archimedes spiral (Fig. 8, lower).
When viewed from the rotating frame, the projectile is obviously deflected to the right, and from
the azimuthal component of Eqn. (31) we readily attribute this to the Coriolis force,

dr0 dr0
2ω 0 M = −2Ω M,
dt dt
since ω 0 = Ω. Notice that the horizontal speed and thus the kinetic energy increase with time (Fig. 9,
lower). The rate of increase of rotating frame kinetic energy (per unit mass) is

dV02/2 d(U02 + r02 Ω2 )/2 dr0 0 2


= = rΩ
dt dt dt
where the term on the right side is the work done by the centrifugal force, r0 Ω2 , on the radial velocity,
dr0 /dt. If the projectile had not returned to the ground, its speed (observed from the rotating reference
frame) would have increased without limit so long as the radius increased.24

3.4 Appendix to Section 3: Cylindrical polar coordinates.

Rotational phenomena are often analyzed most efficiently with cylindrical polar coordinates, reviewed
here briefly. The vertical coordinate is exactly the z or x3 of Cartesian coordinates, and we need
consider only the horizontal (two-dimensional) position, which can be specified by a distance from the
origin, r, and the angle, θ between the radius vector and (arbitrarily) the x1 axis (Fig. 10). The
corresponding unit vectors are given in terms of the time-independent Cartesian unit vectors that define
the stationary frame by

er = cos(θ )e1 + sin(θ )e2 , and, eθ = −sin(θ )e1 + cos(θ )e2 , (41)
24
In the previous example, walking on a merry-go-round, it was suggested that you would be able to feel the Coriolis force
directly. Imagine that you are riding along on this projectile (don’t try this one at home) — would you be able to feel the
Coriolis force?
3 INERTIAL AND NONINERTIAL DESCRIPTIONS OF ELEMENTARY MOTIONS 34

e
θ e
r

Figure 10: The unit vectors e1 , e2


define a stationary reference frame.
P The unit vectors for a polar coor-
e2 dinate system, er and eθ , are de-
fined at the position of a given
X parcel, P. These unit vectors are
time-dependent since the angle θ
is time-dependent.
θ e1

where both r and θ are time-dependent. The position vector in this system is

X = rer ,

and hence the velocity is


dX dr der dr
= er + r = er + rω eθ , (42)
dt dt dt dt
where we have taken account of the time-dependence of er and ω = d θ /dt. Continuing, the equation of
motion is
d 2X
 2  


d r 2 dr
M = − rω Mer + 2ω + r Meθ = Fr er + Fθ eθ . (43)
dt 2 dt 2 dt dt
Sorted by radial and tangential components this is just
 2 
d r 2
− rω M = Fr , (44)
dt 2




dr
2ω + r M = Fθ . (45)
dt dt
Notice that there are terms, −rω 2 and 2ω drdt , that look like centrifugal and Coriolis force terms and are
sometimes deemed to be such, even by otherwise very careful authors, e.g., Boas, p. 399.14 However,
these equations have been written for an inertial reference frame where centrifugal and Coriolis forces
do not arise. Moreover, the angular velocity ω is that of the parcel position, not the reference frame,
and, these terms are part of the directly-observed acceleration seen in the inertial frame.
To find the rotating frame equation of motion is straightforward; the radius is identical in the
rotating frame, r0 = r, since the origins are assumed coincident. The unit vectors are identical since they
are defined at the location of the parcel, e0r = er and eθ0 = eθ ; the components Fr0 = Fr and Fθ0 = Fθ are
thus also identical in the two frames. The only thing different is that the angular velocity ω is
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 35

decomposed into the presumed constant angular velocity of the rotating frame, Ω, and a relative angular
velocity, ω 0 , i.e., ω = Ω + ω 0 . An observer in the rotating reference frame will see the parcel motion
associated with the relative angular velocity, but not the angular velocity of the reference frame, Ω,
though she will know that it is present. Substituting this into the inertial frame equations of motion
above, and rearrangement to put the observed acceleration on the lefthand side (and so moving terms
containing Ω to the right hand side) yields the formidable-looking rotating frame equations of motion,
by radial and tangential components:
 2 0 
d r 0 02
M = r0 Ω2 + 2Ωω 0 r0 M + Fr0 ,

2
−r ω (46)
dt

0 0 dr0
 
0 dr 0 dω
2ω +r M = −2Ω M + Fθ0 . (47)
dt dt dt
There is a genuine centrifugal force term ∝ Ω2 in the radial component (46), and there are Coriolis
force terms, ∝ 2Ω, on the righthand sides of both (46) and (47). We have derived these terms for the
third time now; in Cartesian coordinates, Eqn. (26), in vector form, Eqn. (27), and here in cylindrical
polar coordinates. You should be sure to verify these equations, as they are perhaps the most direct
derivation of the Coriolis force and most easily show where the factor of 2 arises in the Coriolis terms.

4 A reference frame attached to the rotating Earth

The equations of motion appropriate to the atmosphere and ocean differ from the general rotating frame
equations considered up to now in two significant ways. First, it isn’t just the reference frame that
rotates, but the entire Earth, ocean and atmosphere, aside from the comparatively small (but very
important!) relative motion of winds and ocean currents. A consequence is that the horizontal
component of the centrifugal force on a stationary parcel is exactly canceled by a component of the
gravitational mass attraction and the centrifugal force does not appear in the rotating frame equations of
motion for the atmosphere and oceans, a very welcome simplification (Section 4.1). Second, because
the Earth is nearly spherical, Earth’s rotation vector has a significant angle to the local horizontal,
except at the poles. The horizontal component of the Coriolis force thus varies with latitude (Section
4.2, and with consequences considered in Part 3). Lastly, the inertial and rotating frame descriptions of
a very simple geophysical flow is considered in Section 4.3.
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 36

1.2 Figure 11: Cross-section through a hemisphere



ellipse of a gravitating and rotating planet. The grav-
sphere
1 itational acceleration due to mass attraction is
shown as the vector g∗ that points to the cen-
0.8 ter of a spherical, homogeneous planet. The
c centrifugal acceleration, C, associated with the
0.6 planet’s rotation is directed normal to and away
from the rotation axis. The combined gravita-
0.4 g* tional and centrifugal acceleration is shown as
g
the heavier vector, g. This vector is in the di-
0.2 rection of a plumb line, and defines vertical. A
surface that is normal to g similarly defines a
0 level surface, and has the approximate shape
of an oblate spheroid (the solid curve). The el-
−0.2 lipse of this diagram has a flatness F = 0.1 that
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
approximates Saturn; for Earth, F = 0.0033.

4.1 Cancellation of the centrifugal force

4.1.1 Earth’s (slightly chubby) figure

If Earth was a perfect, homogeneous sphere, the gravitational mass attraction at the surface, g∗, would
be directed towards the center (Fig. 11). Because the Earth is rotating, every parcel on the surface is
also subject to a centrifugal force of magnitude Ω2 R sin θ , where REarth is Earth’s nominal radius, and θ
is the colatitude (π /2 - latitude). This centrifugal force has a component parallel to the surface (a shear
stress)
Cθ = Ω2 REarth sin θ cos θ , (48)
that is directed towards the equator (except at the equator where the 3-d vector centrifugal force is
vertical).25 Cθ is not large compared to g∗, Cθ /g∗ ≈ 0.002 at most, but it has been present since the
Earth’s formation. A fluid can not sustain a shear stress without deforming, and over geological time
this holds as well for the Earth’s interior and crust. Thus it is highly plausible that the Earth long ago
settled into an equilibrium configuration in which this Cθ is exactly balanced by a component of the
gravitational (mass) attraction that is parallel to the displaced surface and poleward.
To make what turns out to be a rough estimate of the displaced surface, η , assume that the
gravitational mass attraction remains that of a sphere and that the meridional slope times the
25 Ancient critics of the rotating Earth hypothesis argued that loose objects on a spinning sphere should fly off into space,
which clearly does not happen. Even so, given this persistent centrifugal force, why don’t we drift towards the equator? Alfred
Wegner proposed this as the motive force of Earth’s moving continents (see D. McKenzie, ’Seafloor magnetism and drifting
continents’, in A Century of Nature, 131-137. Ed. by L. Garwin and T. Lincoln, The Univ. of Chicago Press, Chicago, Il,
2003.).
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 37

gravitational mass attraction is in balance with the tangential component of the centrifugal force,
g∗ d η
= Ω2 REarth sin θ cos θ . (49)
REarth d θ
This may then be integrated with latitude to yield the equilibrium displacement
Z θ 2 2
Ω REarth Ω2 R2Earth
η (θ ) = sin θ cos θ d θ = sin θ 2 + constant. (50)
0 g∗ 2g∗
When this displacement is added onto a sphere the result is an oblate (flattened) spheroid, (Fig. 11),
which is consistent qualitatively with the observed shape of the Earth.26 A convenient measure of
flattening is F = (Reqt − R pol )/Reqt , where the subscripts refer to the equatorial and polar radius.
Earth’s flatness is F = 0.0033, which seems quite small, but is nevertheless highly significant in ways
beyond that considered here.27 The flatness of a rotating planet is given roughly by F ≈ Ω2 R/g. If the
gravitational acceleration at the surface, g, is written in terms of the planet’s mean radius, R, and density,
ρ , then F = Ω2 /( 34 π Gρ ), where G = 6.67×10−11 m3 kg−1 s−2 is the universal gravitational constant.
The rotation rate and the density vary a good deal among the planets, and consequently so does F. The
gas giant, Saturn, has a rotation rate a little more than twice that of Earth and a very low mean density,
about one eighth of Earth’s. The result is that Saturn’s flatness is large enough, F ≈ 0.10, that it can be
discerned through a backyard telescope (Fig. 11).

4.1.2 Vertical and level in an accelerating reference frame

Closely related is the notion of ’vertical’. A direct measurement of vertical can be made by means of a
plumb bob; a plumb line that is at rest is parallel to the local gravity and in the direction vertical.
Following the discussion above we know that the time-independent, acceleration field of the Earth is
made up of two contributions, the first and by far the largest being mass attraction, g∗, and the second
being the centrifugal acceleration associated with the Earth’s rotation, C, Fig. (11). Just as on the
merry-go-round, this centrifugal acceleration adds with the gravitational mass attraction to give the net
26 The pole-to-equator rise given by Eqn. (50), is about 11 km. Precise observations show that Earth’s equatorial radius,
Reqt = 6378.2, is greater than the polar radius, R pol = 6356.7 km, by about 21.5 km. The calculation made here is a first
approximation that ignores the gravitational mass attraction of the displaced mass, which is also toward the equator. Thus still
more mass must be displaced in order to achieve a gravitational/rotational equilibrium with the net result being about a factor
of two greater amplitude than Eqn. (50) indicates.
A comprehensive source for physical data on the planets is C. F. Yoder, ‘Astrometric and geodetic data on Earth and the
solar system,’ Ch. 1, pp 1–32, of A Handbook of Physical Constants: Global Earth Physics (Vol. 1). American Geophysical
Union (1995).
27 To note just two: 1) Earth’s ellipsoidal shape must be accounted for in highly precise, long range navigation systems

(GPS), while shorter range or less precise systems can approximate the Earth as spherical. 2) Because the Earth is not
perfectly spherical, the gravitational tug of the Sun, Moon and planets can exert a torque on the Earth and thereby perturb
Earth’s rotation vector.19
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 38

acceleration, called ’gravity’, g = g∗ + C, a time-independent vector (field) whose direction is


observable with a stationary plumb line and whose magnitude made be inferred by observing the period
of an oscillating plumb bob (a simple pendulum). A surface that is normal to the gravitational
acceleration vector is said to be a level surface in as much as the acceleration component parallel to that
surface is zero. A level surface can also be defined by observing the free surface of a water body that is
at rest in the rotating frame, since a resting fluid can sustain only normal stresses, i.e., pressure but not
shear stress. Thus the measurements of vertical and level that we can readily make necessarily lump
together the centrifugal force due to Earth’s rotation with gravitational mass attraction. The happy result
is that the rotating frame equation of motion applied in an Earth-attached reference frame, Eqn. (2),
does not include the centrifugal force associated with Earth’s rotation (and neither do we tend to roll
towards the equator!).

4.1.3 The equation of motion for an Earth-attached frame

The inertial and rotating frame momentum equations are listed again for convenience using velocity in
place of the previous time rate change of position,
dV
M = F + g∗M, and, (51)
dt
dV0
M = −2Ω Ω ×V0 M − Ω ×Ω Ω ×X0 M + F0 + g0∗ M, (52)
dt
and note that the inertial frame velocity of Eqn. (51) may be written in terms of the relative velocity as
V = VΩ + V0 , (53)
where
Ω ×X
VΩ = −Ω
is the planetary velocity due to the solid body rotation of the Earth (Section 2.2). Now we are going to
assume the result from above, that there exists a tangential component of gravitational mass attraction
that exactly balances the centrifugal force due to Earth’s rotation and that we define vertical in terms of
the measurements that we can readily make; thus g = g∗ + Ω ×Ω Ω ×X.30 The equations of motion are
then, for the inertial frame,
dV
M = Ω ×Ω Ω ×X M + F + g M, (54)
dt
and for the rotating frame,
dV0
M = − 2Ω Ω ×V0 M + F0 + g M. (55)
dt
Finally we have come to Eqn. (2), which we now see is the rotating frame equivalent of Eqn. (54) (and
we will return to these equations in Section 4.3).28
28 These rather formal notions of vertical and level turned out to have considerable practical importance beginning on a
sweltering August afternoon when the University Housing Office notified your dear younger brother, Gustave-Gaspard (GG),
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 39

4.2 Coriolis force on motions in a thin, spherical shell

Application to geophysical flows requires one further small elaboration because the rotation vector
makes a considerable angle to the vertical except at the poles. An Earth-attached, rectangular coordinate
system is usually envisioned to have ex in the east direction, ey in the north direction, and horizontal is
defined by a tangent plane to Earth’s surface. The vertical direction, ez , is thus radial with respect to the
(approximately) spherical Earth. The rotation vector Ω thus makes an angle φ with respect to the local
horizontal x0 , y0 plane, where φ is the latitude of the coordinate system and thus

Ω = Ω cos(φ )ey + Ω sin (φ )ez.

If V0 = u0 ex + v0 ey + w0 ez , then the Coriolis force (divided by the mass, M) is

Ω ×V0 = −(2Ω cos (φ )w0 − 2Ω sin (φ )v0 )ex − 2Ω sin (φ )u0 ey + 2Ω cos (φ )u0 ez .
− 2Ω (56)

Large scale geophysical flows are very flat in the sense that the horizontal components of wind or
current are very much larger than the vertical component, u0 ∝ v0  w0 , in part because the oceans and
the atmosphere are quite thin, having a depth to width ratio of about 0.001. The ocean and atmosphere
are also stably stratified in the vertical, which still further inhibits the vertical component of motion. For
large scale (in the horizontal) flows, the Coriolis term multiplying w0 in the x component of Eqn. (56) is
thus very much smaller than the terms multiplied by u0 or v0 and as an excellent approximation the w0
terms may be ignored, very often with no mention. The Coriolis term that appears in the vertical
component is usually much, much smaller than the gravitational acceleration, and it too is almost
always dropped without mention. The result is the thin fluid approximation of the Coriolis force in
which only the horizontal Coriolis force acting on horizontal motions is retained,

Ω ×V0 ≈ −f×V0 = f v0 ex − f u0 ey
−2Ω (57)

The vector f = f ez , where f is the very important Coriolis parameter,

f = 2Ω sin (φ ) (58)

that because of an unexpectedly heavy influx of freshmen, his old and comfortable dorm room was not going to be available.
As a consolation, they offered him the use of the merry-go-round (the one in Section 3.3, and still running) at the local, failed
amusement park just gobbled up by the University. Knowing well you enthusiasm for rotation, he accepts. The centrifugal
force, amusing at first, was soon a huge annoyance. GG suffered from recurring nightmares of sliding out of bed and over a
cliff. Something had to be done, so you decide to build up the floor so that the tilt of the floor, combined with gravitational
acceleration, would be just sufficient to balance the centrifugal force, as in Eqn. (49). What shape η (r) is required, and how
much does the outside edge (r = 6 m, Ω = 0.5 rad s−1 ) have to be built up? How could you verify success? Given that GG’s
bed is 2 m long and flat, what is the axial traction, or tidal force? Is the calibration of a bathroom scale effected? Guests
are always impressed with GG’s rotating dorm room, and to make sure they have the full experience, he sends them to the
refrigerator for another cold drink. Describe what happens next using Eqn. (55). Is their experience route-dependent?
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 40

and φ is the latitude. Notice that f varies with the sine of the latitude, having a zero at the equator and
maxima at the poles; f < 0 in the southern hemisphere.29
For problems that involve parcel displacements, L, that are very small compared to the radius of
the Earth, REarth, a simplification of f is often appropriate. The Coriolis parameter may be expanded in
a Taylor series about a central latitude φ0 where the north coordinate y = y0 ,

df
f (y) = f (y0 ) + (y − y0) |y +HOT. (59)
dy 0

If the second term involving the first derivative d f /dy = 2Ωcos(φ )/REarth, often written as d f /dy = β ,
is demonstrably much smaller than the first term, which follows if L  REarth, then the second and
higher terms may be dropped to leave f = f (y0 ), a constant. Under this so-called f -plane
approximation, the period of inertial motions, 2π / f , is just a little bit less than 12 hrs at the poles, a
little less than 24 hrs at 30 N or S, and infinite at the equator. (Infinite is, of course, unlikely physically,
and suggests that something different will arise on the equator; more on this below in footnote 32). The
period of inertial motions is sometimes said to be half of a ’pendulum day’, the time required for a
Foucault pendulum to precess through 2π radians.30
29 The Coriolis parameter f vanishes at the equator as does the horizontal component of the Coriolis force. However, as we
noted in the discussion of Fig. 3, the three-dimensional (vector) Coriolis force is independent of the location of the velocity
and the rotation vectors. To appreciate the significance of this, consider relative velocities that are eastward and northward,
and sketch the resulting (three-dimensional) Coriolis force vector ∝ −2Ω Ω ×V0 at several latitudes that span pole-to-pole.
Except at the poles, the three-dimensional Coriolis force has a vertical component that is negligible for most atmospheric and
oceanic dynamics, but is very important in the context of ship-based gravimetric studies where it is termed the Eotvos effect,
see http://en.wikipedia.org/wiki/Eotvos effect (you may have to type this into your web browser).
30 The effect of Earth’s rotation on the motion of a simple (one bob) pendulum, called a Foucault pendulum in this context,

is treated in detail in many physics texts, e.g. Marion5 , and need not be repeated here. Here are a few questions, however.
Can you calculate the Foucault pendulum motion by rotating the inertial frame solution for a simple pendulum? How does
the time required for precession through 360 degrees depend upon latitude? What happens when the pendulum’s natural
frequency (in the absence of Earth’s rotation) equals the Earth’s rotation rate? Given the rotated trajectory, can you show that
the acceleration of the bob for very short times is consistent with the rotating frame equations of motion?
Foucault pendulums are commonly displayed in science museums, though seldom to large crowds (but see The Prism and
the Pendulum by R. P. Crease for a more enthusiastic appraisal). It is worthwhile and easy to make and observe your very own
Foucault pendulum, a simple pendulum having two readily engineered properties. First, the e-folding time of the motion due
to frictional dissipation must be long enough that the precession will become apparent before the motion dies away, 20 min
will suffice at mid-latitudes. This can be achieved using a dense, smooth and symmetric bob having a weight of about half a
kilogram or more, and suspended on a fine, smooth monofilament line. It is helpful if line is several meters or more in length.
Second, the pendulum should not interact appreciably with its mounting. This is harder to evaluate, but generally requires a
very rigid support, and a bearing that can not exert torque, for example a fish hook bearing on a very hard steel surface. The
rotation effect on a simple pendulum is proportional to the rotation rate, and so you should plan to bring a simple and rugged
pocket pendulum (a rock on a string will do) on your merry-go-round ride (Section 3.2). How do your observations (even if
qualitative) compare with your solution for a Foucault pendulum? (Hint - consider the initial condition.)
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 41

4.3 Why do we insist on the rotating frame equations?

We have emphasized that the rotating frame equation of motion has some inherent awkwardness, viz.,
the loss of Galilean invariance and global momentum conservation. Why, then, do we insist upon using
the rotating frame equations for nearly all of our analyses of geophysical flow? The reasons are several,
any one of which would be compelling, but beginning with the fact that the definition and
implementation of an inertial frame (outside of the Earth) is simply not a viable option; whatever
simplicity we might gain by omitting the Coriolis force would be lost to difficulty with observation.
Consider just one aspect of this: the inertial frame (or absolute) velocity (Eqn. 53), V = VΩ + V0 , is
dominated by the planetary velocity due to the solid-body rotation VΩ = ΩRe cos(latitude), where REarth
is earth’s nominal radius, 6365 km, and thus VΩ ≈ 400 m s−1 near the equator. By comparison, a very
fast wind speed at mid-level of the atmosphere is V 0 ≈ 50 m s−1 (the westerlies of Fig. 1) and a
comparable ocean current is V 0 ≈ 2 m s−1 (the western boundary current of Fig. 2). The very large
planetary velocity VΩ is accelerated centripetally by a tangential (almost) component of gravitational
mass attraction associated with the ellipsoidal shape of the Earth discussed in Section 4.1 that is larger
than the Coriolis force in the ratio VΩ/V 0 that is O(10) for the atmosphere, and much more for ocean
currents.31 The inertial frame equations have to account for VΩ and this very large centripetal force
explicitly, and yet our interest is almost always the small relative motion of the atmosphere and ocean,
V0 , since it is the relative motion that transports heat and mass over the Earth. In that important regard,
the planetary velocity VΩ = is invisible to us Earth-bound observers, no matter how large it is. To say it
a little differently — it is the relative velocity that we measure when observe from Earth’s surface, and
it is the relative velocity that we seek for almost every practical purposes — the Coriolis force follows.

4.3.1 Inertial oscillations from an inertial frame

Given that our goal is the relative velocity, then the rotating frame equation of motion is generally much
simpler and more appropriate than is the inertial frame equation of motion. To help make this point we
will analyze the free oscillations of Eqs. (54) and (55), i.e., F = F0 = 0, usually called inertial
oscillations, that are interesting in their own right. The domain is presumed to be a small region
centered on the pole so that latitude = 90 degrees, and the domain is, in effect, flat. We will consider
only horizontal motions, and assume the realistic condition that the motion will be a small perturbation
away from the solid body rotation, V0  VΩ . The motion viewed from the inertial frame is thus almost
circular and it is appropriate to use the cylindrical coordinate momentum equation, Eqn. (31) (dividing
31 How does the slope of the 500 mb pressure surface at mid-latitudes (Fig. 1) compare with the flatness of Earth?
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 42

out the constant M):

d 2r
− rω 2 = −Ω2 r, (60)
dt 2
dr dω
2ω + r = 0. (61)
dt dt
Notice that when ω = Ω and dr/dt = 0, the motion is balanced in the sense that d 2r/dt 2 = 0 and r
remains constant. We are going to assume an initial condition that is a small radial perturbation away
from such a balanced state, r = Ro and ω = Ω. Since there is no tangential force, Eqn. (61) may be
integrated,
ω r2 = ΩR2o = A, (62)
which shows that the angular momentum, A, is a conserved quantity. Eqn. (62) can then be used to
eliminate ω from Eqn. (60) to yield an equation for r(t) alone,

d 2r A2
− = −Ω2 r. (63)
dt 2 r3

To solve this equation it is convenient to move the centripetal acceleration term A2 /r3 to the right
side where it will be summed with the centripetal force,

d 2 r A2
= 3 − Ω2 r, (64)
dt 2 r
yielding what looks just like an equation of motion. However, it is important to understand that d 2r/dt 2
is not the radial component of acceleration that is observed in either an inertial or a rotating reference
frame (cf. Eqs. 31 and 31) and to acknowledge this explicitly, the right hand side of Eqn.(64) will be
called a pseudo (or false) force. None of this effects the solution per se, but only the words we use and
the inferences we might then draw. And specifically, if you measured the radial force/M on the parcel
you wouldn’t find A2 /r3 − Ω2 r, but rather −Ω2r, the right hand side of Eqn. (63).
Eqn. (64) is a well-known, nonlinear oscillator equation and is not difficult to solve. However,
because our interest is in the case of small displacements away from the balanced state, r = Ro , a
simplification is appropriate. To clarify what is meant by small displacement it is helpful to write the
radius as
r(t) = Ro (1 + δ (t)).
The meaning of ’small displacement’ is that δ should be small compared to 1. Substitution into Eqn.
(64) and rearrangement yields

d 2δ
= Ω2 (1 + δ )−3 − Ω2 (1 + δ ). (65)
dt 2
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 43

1.5

1 −3
(1 + δ)
Figure 12: The terms of the
2
pseudo force/Ω

0.5
−3
(1 + δ) − (1 + δ) right side of Eqn. (65), dubbed
pseudo forces and normalized by
0 Ω2 , shown as a function of δ . Note
that the net pseudo force (solid
−0.5
linearized, −4δ
line) is nearly linear in δ when δ
−(1 + δ) is small, roughly δ ≤ 0.1.
−1

−1.5
−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2
δ

When we plot the right side of Eqn. (65) it is evident that the net pseudo force is a nearly linear function
of δ provided δ ≤ 0.1. To exploit this we can expand the nonlinear term of Eqn.(65) in Taylor series
about δ = 0,

d 2δ
= Ω2 (1 − 3δ + 6δ 2 + HOT ) − Ω2(1 + δ )
dt 2
≈ −4Ω2 δ , (66)

where HOT are terms that are higher order in δ . For small displacements the quadratic and higher order
terms may be neglected, leaving a simple harmonic oscillator equation, Eqn. (66), at a frequency 2Ω.
If the initial condition is a radial impulse that gives a radial velocity V0, then the initial condition
for Eqn. (66) is

(t = 0) = (Vo/Ro ) cos2Ωt.
dt
The solution for δ is
δ (t) = (V0 /2Ω) sin (2Ωt)
and the radius is then
r(t) = Ro (1 + δ0 sin (2Ωt)), (67)
where δ0 = V0/2Ω. The corresponding angular rotation rate can be found by using Eqn. (67) together
with the angular momentum conservation Eqn. (62),

ω (t) = ≈ Ω(1 − 2δ0 sin (2Ωt)). (68)
(1 + δ0 sin (2Ωt))2
When graphed, these show that the parcel moves in an ellipsoidal orbit, Fig. (13, left panels), that
crosses the (balanced) radius r = Ro four times per complete orbit. The rotating frame turns through
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 44

180 degrees just as the parcel returns to r = Ro the second time, after completing a full cycle of the
oscillation. When viewed from the rotating frame (Fig. 13, right panels), the parcel appears to be
moving in a clockwise-orbiting, circular path, with a frequency 2Ω.

4.3.2 Inertial oscillations from the rotating frame

It is convenient to expand the rotating frame equation of motion (55) in Cartesian coordinates. Since we
have restricted the analysis above to small displacements we can utilize the f-plane approximation that
takes f as a constant. The accelerations of the horizontal components u0 , v0 are then

du0 dv0
= f v0 and = − f u0 . (69)
dt dt
Given that the initial condition is an impulse causing a small velocity V0 in the y-direction then the
solution for velocity and displacement is just

u0 = V0 sin ( f t), and v0 = V0 cos ( f t), (70)


0 0
x = δ0(1 − cos ( f t)), and y = δ0 sin ( f t), (71)

where the displacement is δ0 = V0/ f . The velocity of the parcel seen from the rotating frame, V0 ,
rotates at a rate of f = 2Ω in a direction opposite the rotation of the reference frame, Ω.32 This is
exactly the result found in the inertial frame analysis but was far simpler to obtain because we did not
have to account for the absolute velocity, V = VΩ + V0 , but the relative velocity only. From the rotating
frame perspective, Eqn. (55), the rotation of the velocity vector is attributable to deflection by the
32 Three questions for you: 1) How does this compare with the momentum balance and motion described by Eqn. (32)?
2) Suppose that the impulse was in the azimuthal direction, what would change? 3) Notice that the parcel displacement δ0
associated with an inertial oscillation goes as 1/ f , and hence for a given impulse, δ0 → ∞ as f → 0, i.e., as the latitude
approaches the equator. We can be pretty sure that something will intervene to preclude infinite displacements at the equator.
One possibility is that the north-south variation of f around the equator will become relevant as the displacement becomes
large, i.e., the f −plane assumption that δ0  REarth noted with Eqn. (59) will break down. Suppose instead that we keep the
first order term in f (y), and assume f = β y, an equatorial beta-plane. Describe the equatorial inertial oscillations of a parcel
initially on the equator, and given an impulse V0 directed toward the northeast. How about an impulse directed toward the
northwest? You should find that these two cases will yield quite different trajectories. This is an example, of which we will
see more in Part 2, of the anisotropy that arises from rotation and Earth’s spherical shape.
4 A REFERENCE FRAME ATTACHED TO THE ROTATING EARTH 45

an inertial frame (blue) and a rotating frame (red) as observed in the rotating frame (red)
and as computed in the rotating frame (green)
dV/dt = −g ∇ Φ

V(0) = V dV‘/dt = − 2 Ω x V‘
0

V‘(0) = V0 − Ω x X‘

V‘
0

1 1
Y‘ Y‘

0.5 1.5 0.5 1.5


Z

Z
1 1
0 X‘ 0.5 0 X‘ 0.5
−1.5 −1.5
0 0
−1 −1
−0.5 −0.5 −0.5 −0.5
0 0
0.5 −1 0.5 −1
time/(2 π/Ω) = 0 1 Y time/(2 π/Ω) = 0 1 Y‘
1.5 −1.5 1.5 −1.5
X X‘

an inertial frame (blue) and a rotating frame (red) as observed in the rotating frame (red)
and as computed in the rotating frame (green)
dV/dt = −g ∇ Φ

V(0) = V dV‘/dt = − 2 Ω x V‘
0

V‘(0) = V − Ω x X‘
0

1 1
X‘ Y‘

0.5 Y‘ 1.5 0.5 1.5


Z

1 1
0 0.5 0 X‘ 0.5
−1.5 −1.5
0 0
−1 −1
−0.5 −0.5 −0.5 −0.5
0 0
0.5 −1 0.5 −1
time/(2 π/Ω) = 0.27 1 Y time/(2 π/Ω) = 0.27 1 Y‘
1.5 −1.5 1.5 −1.5
X X‘

an inertial frame (blue) and a rotating frame (red) as observed in the rotating frame (red)
and as computed in the rotating frame (green)
dV/dt = −g ∇ Φ

V(0) = V dV‘/dt = − 2 Ω x V‘
0

V‘(0) = V − Ω x X‘
0

1 1
Y‘

0.5 X‘ 1.5 0.5 1.5


Z

1 1
0 0.5 0 X‘ 0.5
−1.5 Y‘ −1.5
0 0
−1 −1
−0.5 −0.5 −0.5 −0.5
0 0
0.5 −1 0.5 −1
time/(2 π/Ω) = 0.51 1 Y time/(2 π/Ω) = 0.51 1 Y‘
1.5 −1.5 1.5 −1.5
X X‘

Figure 13: The two-dimensional trajectory of a parcel subject to a centripetal force, −rΩ2 , as if it were
on a frictionless parabolic surface. The initial velocity was a solid body rotation in balance with the
centripetal force, and a small radial impulse was then superimposed. In this case the ratio V 0 /VΩ ≈ 0.2,
which is far larger than actually occurs in the ocean or atmosphere. The left column shows the resulting
ellipsoidal trajectory as seen from an inertial frame, along with the circular trajectory that is seen from
a rotating frame (indicated by the rotating, solid unit vectors). The right column shows the trajectory
as seen from the rotating frame only, along with the solution computed in the rotating frame (shown as
green dots). These lie exactly on top of the ’observed’ trajectory and are very difficult to discern if color
is not displayed; try the Matlab script rotation 2.m (Section 7) that includes this and a number of other
cases.
5 A DENSE PARCEL ON A ROTATING SLOPE 46

Coriolis force.33 This kind of motion, termed an inertial oscillation,34 is frequently observed in the
upper ocean following a sudden shift in the wind speed or direction (Fig. 14).

5 A dense parcel on a rotating slope

The motion of a dense parcel released onto a sloping, rotating sea floor illustrates several important
consequences of rotation. The sea floor is presumed to have a depth z = −b(y) that increases in the y
direction with slope db/dy = α , a constant. The buoyancy of the parcel is g0 = −g δρρo , where δ ρ is the
density anomaly of the parcel with respect to its surroundings (assumed constant), and ρo is a nominal
sea water density, 1030 kg m−3 .35 The component of the buoyancy parallel to the sea floor, g0 α , thus
provides a constant force (per unit mass, understood from here on) in the y direction. The effects of
rotation will be modeled by the thin fluid form of the Coriolis force and the Coriolis parameter f will be
assumed constant. If the parcel is in contact with a sloping bottom, then it is plausible that the
momentum balance should include a representation of bottom friction. The bottom friction will be
represented by the simplest linear (or Rayleigh) law in which the friction is presumed to be proportional
to and antiparallel to the velocity difference between the current and the bottom, i.e., bottom friction
= −r(V − Vbot ).36 The ocean bottom is at rest in the rotating frame and hence Vbot = 0 and omitted
33 We noted in Section 3.4 that the rotating frame equations of motion does not support global momentum conservation or
Galilean invariance. The former can be seen by noting that if all forces except Coriolis were zero, and the initial condition
included a velocity, then that velocity would be continually deflected and change direction (as an inertial oscillation) with
nothing else showing a reaction force; i.e., global momentum would not be conserved. This evident nonconservation is
ignorable in most practical analyses.
The Coriolis force is isomorphic to the Lorentz force, qV×B, on a moving, charged particle in a magnetic field B. Thus a
charged particle moving through a uniform magnetic field will be deflected into a circular orbit with the cyclotron frequency,
qB/M, analogous to an inertial oscillation at the frequency f . General Relativity predicts that a rotating object is accompanied
by a gravitomagnetic field that gives rise to a Coriolis-like force on moving objects. The Gravity Probe B mission, one of the
most challenging experiments ever conducted, has apparently confirmed the presence of a gravitomagnetic field around Earth,
see http://einstein.stanford.edu/
34 The name ‘inertial oscillation’ is very widely accepted but is not highly descriptive of the dynamics in either the rotating

or inertial reference frame. For the rotating frame, ‘Coriolis oscillation’ might be more appropriate, and for the inertial frame
see D. R. Durran, ‘Is the Coriolis force really responsible for the inertial oscillation?’ Bull. Am. Met. Soc., 74(11), 2179–2184
(1993).
35 Notice that a prime superscript is used here to denote reduced gravity. The prime previously used to indicate rotating

frame velocity will be omitted, with rotating frame understood.


36 This use of a linear friction law is purely expedient. A linear friction law is most appropriate in a viscous, laminar

boundary layer that is in contact with a no-slip boundary. In that case τ = µ ∂∂Uz within the laminar boundary layer, where
µ is the viscosity of the fluid. However, the laminar boundary layer above a rough ocean bottom is very thin, O(10−3) m,
and above this the flow will in general be turbulent. If the velocity that is used to estimate or compute friction is measured
or computed within the much thicker turbulent boundary layer, as it almost always has to be, then the friction law is likely
better approximated as independent of the viscosity and quadratic in the velocity, τ = ρCd U 2 , where Cd is the drag coefficient.
Typically, Cd = 1 − 2 × 10−3, depending upon bottom roughness, mean speed, and more.
5 A DENSE PARCEL ON A ROTATING SLOPE 47

current at 25 m depth; passage of Felix


2
Felix east
speed, m s−1

1 north

0
a
−1
0 1 2 3 4 5 6 7 8 9 10

1
speed, m s−1

0.5
0
−0.5 b

0 1 2 3 4 5 6 7 8 9 10
−5
x 10
5
accel, m sec−2

east
north
0
c
−5
0 1 2 3 4 5 6 7 8 9 10
time/(2π/f)

Figure 14: Ocean currents at a depth of 25 m, measured by a current meter deployed southwest of
Bermuda. The time scale is inertial periods, 2π / f , which are nearly equal to days at this latitude. Hur-
ricane Felix passed over the current meter mooring at the time noted at upper left, and the strong and
rapidly changing wind stress produced energetic, clockwise rotating currents within the upper ocean. To
a first approximation these are inertial oscillations. They differ from pure inertial oscillations in that
their frequency is roughly 5% percent higher than f and their amplitude e-folds over about 10 days (by
inspection of these data). These small departures from pure inertial motion are indicative of wave-like
dynamics considered in Part 2. (a) East and north current components. (b) Current vectors, with north
’up’. (c) Acceleration estimated from the current meter data as dV0 /dt + 2Ω
Ω ×V0, as if the measurements
were made on a specific parcel. The large acceleration to the west northwest corresponds in time to the
passage of Felix. The direction of the estimated acceleration is roughly parallel to the observed winds
(not shown here). Notice the much smaller oscillations having a period of about 0.5 inertial periods (es-
pecially for t > 8). These are very likely due to pressure gradients associated with the semidiurnal tide.
This is a small part of the data described in detail by Zedler, S.E., T.D. Dickey, S.C. Doney, J.F. Price,
X. Yu, and G.L. Mellor, ’Analysis and simulations of the upper ocean’s response to Hurricane Felix at
the Bermuda Testbed Mooring site: August 13-23, 1995’, J. Geophys. Res., 107, (C12), 25-1 - 25-29,
(2002), available online at http://www.opl.ucsb.edu/tommy/pubs/SarahFelixJGR.pdf.
5 A DENSE PARCEL ON A ROTATING SLOPE 48

from here on. From observations of ocean currents a reasonable value of the bottom friction coefficient
r for a density-driven current on a continental shelf is r = O(10−5) s−1 which is roughly an order of
magnitude smaller than a mid-latitude value of f . Since r appears in the momentum equations in the
same way that f does, it is expected that frictional effects will be small compared with rotational effects.
The equations of motion including bottom friction are:
du
= f v − ru, (72)
dt
dv
= − f u − rv + g0 α ,
dt
with vector equivalent,
dV
= − f k×V − rV + g0 ∇b (73)
dt
Initial conditions are a state of rest,

u(0) = 0 and v(0) = 0 or, V(0) = 0

The solution of this linear model,


g0 α
u(t) = [ f − exp (−rt)( f cos ( f t) − r sin ( f t))] , (74)
r2 + f 2
g0 α
v(t) = 2 [r − exp (−rt)( f sin ( f t) + r cos ( f t))] ,
r + f2
is not complex by the standards of fluid dynamics, but it does contain three parameters along with the
time, and so has a fairly large parameter space. Our interest is not any one specific solution but rather to
gain some understanding of the quantitative effects of rotation and friction over the entire family of
solutions. How can the solution be displayed to this end?
One approach that is very widely applicable is to rewrite the governing equations and (or) the
solution using nondimensional variables. This will serve to reduce the number of parameters to the
fewest possible while retaining everything that was present in the dimensional equations. To define
these nondimensional variables we begin by noting that there are three external parameters in the
problem: 1) the buoyancy and bottom slope, g0 α , which always occur in this combination and so count
as one parameter, an acceleration, 2) the Coriolis parameter, f , an inverse time scale, and 3) the bottom
friction coefficient, r, also an inverse time scale. The inverse Coriolis parameter is used to scale time,
t∗ = t f , since rotational effects will dominate frictional effects in most cases of interest. A velocity
scale is estimated as the product of the acceleration and the rotational time scale, Ugeo = (g0 α )/ f , and
thus the nondimensional velocity is u∗ = u/(g0 α / f ) = u/Ugeo and similarly for the v component; the
subscript geo(strophic) will be discussed shortly. Rewriting the governing equations in terms of these
5 A DENSE PARCEL ON A ROTATING SLOPE 49

particle on a slope w/ rotation and friction, E = 0, 0.05, 0.25

t/(2π/f) = 9.39

5
0
z/(g‘α /f )
2 2

4
−5
3

0 2
10
20 2
30 1 y/(g‘α/f )
40
50
2 60
x/(g‘α/f ) 70 0
80

Figure 15: Dense parcels released onto a rotating slope. The buoyancy force is toward positive y (up
in this figure). (left) Trjectories of three parcels differ by having rather large friction (blue trajectory,
E = 0.25), moderate, more or less realistic friction (green trajectory, E = 0.05) or no friction at all (red
trajectory, E = 0). The elapsed time in units of inertial periods, 2π / f , is at upper left. The along- and
across-slope distance scales are distorted by a factor of almost 10 in this plot, so that the blue trajectory
having E = 0.25 makes a shallower descent of the slope than first appears here. Notice that for values
of E  1 (red and green trajectories), the motion includes a looping inertial motion, and a long-term
displacement that is nearly at right angles to the imposed force, analogous to a geostrophic balance. In
this northern hemisphere ( f ≥ 0) problem in which the force is due to buoyancy and a sloping bottom,
the long term motion is such that shallower bottom depth is to the right when looking in the direction of
the motion. (right) The force balance (solid arrows) and the time-mean motion (the dashed vector) for
the case E = 0.25. The angle of the velocity with respect to the isobaths is E = r/ f , the Ekman number.
The Coriolis force (/M) is labeled −f×V where f is f times a vertical unit vector. Simulations of this
sort are carried out by the script partslope.m (linked in Sec. 7).
5 A DENSE PARCEL ON A ROTATING SLOPE 50

nondimensional variables gives


du∗
= v∗ − Eu∗, (75)
dt∗
dv∗
= −u∗ − Ev∗ + 1,
dt∗
where E is the Ekman number,
r
E=
f
the nondimensional ratio of the friction parameter to the Coriolis parameter (more on E below). The
initial conditions are u∗ (0) = 0, v∗ (0) = 0. The solution to these equations is
1
u∗(t∗ ) = [1 − exp (−Et∗)(cos (t∗ ) − E sin (t∗ ))] , (76)
1 + E2
1
v∗ (t∗) = [E − exp (−Et∗)(sin (t∗ ) + E cos (t∗ ))] ,
1 + E2
and for completeness;
u v
u∗ = , v∗ = , and t∗ = t f .
g0 α / f g0 α / f
The geostrophic velocity scale Ugeo = g0 α / f appears with the velocity only, and thus the single
nondimensional parameter E serves to define the parameter space of this problem. There are other
forms of the Ekman number that follow from different forms of friction parameterization. They all have
in common that small E indicates small friction compared to rotation.37 Trajectories computed from
Eqn. (76) for several values of E are in Fig. (15, left). Trajectories having larger E show both a more
rapid damping of the inertial oscillations and a steeper descent of the slope, i.e., greater frictional
effects. Thus in place of ’large friction’ or ’large rotation’, we have instead large or small E, which
takes account of both r and f .
For a given r it is expected that frictional effects should be greater at lower latitudes. Very near the
equator, E will thus be very large for almost any r, and on that basis alone geostrophic motion would
not be expected near the equator. This result at first appears to be an explanation of both the
ageostrophic pressure-velocity relationship (wind may be normal to pressure isolines and thus parallel
to the pressure gradient) and the comparatively weak pressure gradients that you may have noticed are
characteristic of the velocity and pressure fields near the equator (Fig. 1). Friction may be relevant in
this regard, but a more comprehensive fluid model treated in Part 3 Sec. 3 shows that gravity wave
dynamics is likely to be more important than is friction alone.
37 If this is one of your first encounters with dimensional analysis then this procedure is likely to seem arbitrary and
esoteric. The benefits of dimensional analysis will become clearer with greater experience, mainly, and an attempt to help is
‘Dimensional analysis of models and data sets’, by J. Price, Am. J. Phys., 71(5), 437–447 (2003) and available online in an
expanded version linked in footnote 11. Any specific dimensional analysis may indeed be somewhat arbitrary, as there are
usually several plausible forms. For example, in this problem 1/r coould be used to measure (or nondimensionalize or scale)
the time. How would this change the solution, Eqn. (76)?
5 A DENSE PARCEL ON A ROTATING SLOPE 51

ctd section D
CTD CTD CTD CTD CTD CTD CTD CTD CTD CTD
33 34 35 36 38 39 40 41 42 43
0 27 −500
27.1
−550
200 27.2
−600
27.3
−650
400 27.4
−700
27.5
Depth (m)

depth, m
Density
600 27.6 −750

27.7 −800

800 27.8 −850


27.9
−900
1000 28
−950
28.1
0.5
−1000 0
1200 28.2 −1 −0.5
0 20 40 60 −0.5 0
Distance (km) V, m s−1
U, m s−1

Figure 16: Observations of a dense bottom current on the southern flank of the Scotland-Iceland Ridge.
(left) A section made across the current showing dense water that has come through the narrow Faroe
Bank Channel (about 15 km width, at latitude 62 N and about 90 km to the northeast of this site) and
that will eventually settle into the deep North Atlantic. The units of density are kg m−3, and 1000 has
been subtracted away. (right) A current profile measured at the thick vertical line shown on the density
section. The density section was aligned normal to the isobaths and the current appeared to be flowing
roughly along the isobaths. The core of the dense water has descended roughly 200 m between this site
and the Faroe Bank Channel. Assuming that the descent of the dense water is due mainly to bottom
friction, which trajectory of Fig. (15) is analogous to this current? Said a little differently, what is the
approximate Ekman number of this current?

5.1 Inertial and geostrophic motion

The solution Eqn. (76) can be envisioned as the sum of two distinct kinds of motion, a time-dependent
inertial oscillation that is damped by friction,
   
u∗ exp(−Et∗) cos(t∗) − Esin(t∗ )
= − , (77)
v∗ 1 + E2 sin(t∗ ) + Ecos(t∗ )
and a steady motion that is the single parcel equivalent of geostrophic motion
   
u∗ 1 1
= , (78)
v∗ 1 + E2 E
also modified by friction. In this linear model the amplitude of either kind of motion is directly
proportional to the velocity scale, Ugeo = g0 α / f . For a dense water parcel on a continental slope (Fig.
16) rough values are g0 = g(δ ρ )/ρ0 ≈ g0.5/1000 = 0.5 × 10−2 m s−2 , α = 1.3×10−2 , and f at latitude
62 N is 1.3 ×10−4 s−1 . This gives a geostrophic speed Ugeo ≈ 0.5 m s−1 , which is roughly consistent
with the observed current in Fig. (16).
5 A DENSE PARCEL ON A ROTATING SLOPE 52

The inertial oscillations are the consequence of starting from a state of rest and the imposition of
what amounts to an impulsively started external force (the buoyancy force). This is not an especially
realistic model of a density current released onto a continental slope, but it does serve as a first
approximation of the rapidly changing wind stress exerted upon the upper ocean during the passage of a
storm (Fig. 14). A pure inertial oscillation, that is, an exact two term balance between the time rate of
change and the Coriolis force (Eqn. 69), would not evolve with time; the amplitude would remain
constant and the frequency = f for all time. The amplitude of the (near) inertial oscillations seen in the
field data of Fig. (14) and the model solution of (15) decrease with time; the model-computed
oscillations decay as exp (−Et∗) on account of bottom friction. Given the freedom to adjust r, a
reasonable looking fit could be made to most field data. Whether the physical mechanism of the
observed inertial motion decay is modelled well by linear friction is doubtful regardless of the fit, since
gravity wave dynamics are missing from this model.38
The steady solution (78) is extremely important, as it represent the single parcel equivalent of
damped, geostrophic motion. In a more general vector form, good for any steady, horizontal force G,
and in the absence of friction,
1
Vgeo = − k×G (79)
ρo f
where k is the vertical unit vector. In practice we usually reserve the distinction ’geostrophic’ for the
case that the force is a horizontal pressure gradient, G = −∇P. If the force is the vertical divergence of
a horizontal wind stress, G = ∂ τ /∂ z, then the steady velocity is often called an Ekman velocity. The
key point is that when rotation (the Coriolis force) is present, there can exist a steady velocity that is
perpendicular, or transverse, to an applied force. This solution suggests that a near-geostrophic balance
will result provided that the Ekman number is not too large, say E ≤ 0.1, which commonly occurs, and
provided also that the applied force varies slowly compared to the rotation time scale, 1/ f . This latter
condition holds exactly in these experiments since the bottom slope is spatially uniform and unlimited
in extent.39 The more realistic shallow water (fluid) model of Part 2 will replace this latter condition
with the requirement that the horizontal scale L of a layer thickness (mass) anomaly must exceed the
rotation length scale, C/ f , where C is the gravity wave speed dependent upon stratification.
38 A couple of questions for you: 1) Can you devise an initial condition for this problem that would eliminate the inertial
oscillations? Could you make the inertial oscillations larger than those shown in Fig. (15, left)? To test your hypothesis you
might try experimenting with the Matlab script partslope.m 2) Draw the vector force balance for inertial oscillations (include
the acceleration) with and without bottom friction as in Fig. (15, right). 3) We have pointed out that the value of the linear
friction coefficient r is not well known a priori and so we might regard r as an adjustable parameter. What value of r is
required to mimic the observed decay of near-inertial oscillations of Fig. 16? Does the same model solution account also for
the small, super-inertial frequency shift noted in the field data?
39 A subtle distinction between the present single parcel model result and the usual idea of geostrophy is that the single

parcel model is Lagrangian, i.e., the dependent variables are the position and velocity of a specific parcel, not the velocity at
a point in space as is implicit in Fig. 1, say, which is an Eulerian representation. These two representations merge only in the
highly idealized case that all fields are spatially uniform.
5 A DENSE PARCEL ON A ROTATING SLOPE 53

It is most important to note two properties of the geostrophic velocity of Eqn. (79). First, the
magnitude of Vgeo is inversely proportional to f . Thus for a given G, the geostrophic wind or current
will be larger at a lower latitude. Second, Vgeo is at right angles to the applied force, G, and in the
northern hemisphere, the force acts from right to left, looking in the direction of Vgeo (Fig. 15, right,
where G is g0 α ).

5.2 Energy balance

Energy balance will often be a compact and useful diagnostic. In this problem, the energy source is the
potential energy associated with the dense parcel sitting on a sloping bottom. If the parcel descends the
slope, it will release potential energy and so generate kinetic energy and thus motion.
To find the energy balance equation, multiply the x-component momentum equation by u∗ and the
y-component equation by v∗ and add:

d(u2∗ + v2∗)/2
= v∗ − E(u2∗ + v2∗). (80)
dt∗
Notice that the Coriolis force per se drops out of the energy balance since it is normal to the current and
does no work. The term at left is time rate change of kinetic energy. The first term on the right is the
rate of work by the buoyancy force, which is also the rate of release of potential energy,
PE = g0 (z − z0 ) = g0 α (y − y0 ),

vf dz 0 f2 −dPE 1
v∗ = 0 2
= g α 0 2
= 2
,
(g α ) dt (g α ) dt Ugeo

where the reference depth, z0 , is the initial depth. The second term on the right of (80) is the rate of
work by bottom friction, always negative since bottom friction opposes the velocity. It can be helpful to
integrate (80) with time to compute the change in energy from the initial state:
Rt Rt
(u2∗ + v2∗)/2 − o v∗ dt∗ = − 2 2
o E(u∗ + v∗ )dt∗ ,
(81)
KE + PE = FW,

where KE is the kinetic energy, PE is the change in potential energy as the parcel is displaced up and
down the slope, and FW is the net frictional work done by the parcel, always a loss (Fig. 17).
Immediately after the parcel is released from rest it starts to descend the slope, converting potential
energy into kinetic energy, though with some loss to friction (if E is nonzero). The Coriolis force acts to
deflect the moving parcel to the right, and by about t = 1/ f or t∗ = 1/(2π ), the parcel has been turned
by 1 radian, or about 50◦, with respect to the buoyancy force. The time required for the Coriolis force to
have an appreciable effect on a moving object is thus 1/ f , the very important rotational time scale noted
6 SUMMARY 54

previously. The Coriolis force continues to turn the parcel to the right, and by about t∗ = 1/2 the
velocity is directed up the slope. Work against buoyancy then converts kinetic energy back into potential
energy. The parcel of Fig. (17) does not quite climb the slope to it’s initial height because of frictional
losses, i.e., the sum of potential and kinetic energy is not conserved for parcels having nonzero E.
The Coriolis force does no work on the parcel since it is perpendicular to the velocity, and hence
does not appear directly in the energy balance. The Coriolis force (rotation) nevertheless has a profound
effect on the energy balance overall in as much as it inhibits the release of potential energy. If there is
some friction, as there is in the case shown, then the cross-isobath component of the motion (which
carries the parcel to greater bottom depth and thus releases potential energy) is proportional to the
Ekman number, from Eqn. (76), v∗ /u∗ = E, and thus inversely proportional to f for a given frictional
coefficient, r. Whether friction or rotation is dominant and thus whether the motion is rapidly dissipated
or long-lived, depends solely upon the Ekman number in this highly simplified system (Fig. 15b). In the
important limit that E → 0, and absent inertial motion, the parcel velocity will be perpendicular to the
buoyancy force, as in Eqn. (79), and the parcel will coast along an isobath in steady, geostrophic
motion.40

6 Summary

Section 1 posed the innocent question, What is the Coriolis force? and it is time to summarize an
answer.
1) The origin of the Coriolis force is found in kinematics rather than physics. The Coriolis (and
centrifugal) force is an inertial force that arises from the rotational acceleration of a reference frame
rather than from an interaction between physical objects. Since there is no physical origin for the
Coriolis force, neither should we expect a physical explanation of its origin (until a physical
understanding of inertia itself is at hand18 ). The sole, reliable explanation of the Coriolis force is the
40 Some questions for you: (1) Assuming small Ekman number, how long does it take for a geostrophic balance to arise after
a parcel is released? (2) Are the time-averaged solutions of the single parcel model the solutions of the time-averaged model
equations? Suppose the model equations were not linear, say that friction is ∝ U 2 , then what? (3) Inertial oscillations do not
contribute to the long-term displacement of the parcel, though they can dominate the instantaneous velocity. Can you find an
initial condition on the parcel velocity that prevents these pesky inertial oscillations? You can test your ideas against solutions
from partslope.m (Section 7). (4) Explain in words why a geostrophic balance (or a near geostrophic balance) is expected
in this problem, given only small enough E and sufficient space and time. (5) Make a semi-quantitative test of geostrophic
balance for the westerly wind belt seen in Fig. (1). (6) Sample (by eye) the sea surface height of Fig. (2) along an east-west
section at 33 o N, including at least a few points in the western boundary region. Then estimate the east-west profile of the
inferred geostrophic current (and note that the buoyancy of the sea surface is effectively the full g since the density difference
is between water and air). What is the current direction? Using this result as a guide, sketch the (approximate) large-scale
pattern of surface geostrophic current over the subpolar gyre and lower subtropics on Fig. (2). You can check your result
against formally estimated geostrophic surface currents at http://argo.colorado.edu/∼realtime/welcome/ — you may have to
type this into your web browser.
6 SUMMARY 55
Figure 17: The energy balance for the
trajectory of Fig. (15) having E =
0.2. These data are plotted in a nondi-
2
mensional form in which the energy or
1 work is normalized by the square of the
velocity scale, Ugeo = g0 α / f , and time
energy/(g‘α/f)2

0 is nondimensionalized by the inertial


period, 2π / f . Potential energy was as-
−1 KE signed a zero at the initial depth of the
PE parcel. Note the complementary iner-
−2 KE + PE
FW
tial oscillations of PE and KE, and that
−3 the decrease of total energy was due to
0 1 2 3 4 5 6 7 8 9 10
t/(2π/f)
work against bottom friction (the solid
green and dashed red lines that overlay
one another).

transformation law for acceleration (Section 2.5).


Whether the Coriolis term should be called a force (as done here) or an acceleration is less clear.
The latter is sensible insofar as the Coriolis term arises from the transformation of acceleration, and too
because the Coriolis force on a parcel gives an acceleration that is independent of the mass of the parcel.
However, to understand the contribution of the Coriolis term to the acceleration that is directly observed
in a rotating frame, it is natural to regard the Coriolis term as the Coriolis force.
The choice of a reference frame determines the existence or not of the Coriolis force; either it’s
inertial, in which case there is no Coriolis or centrifugal force, or it’s rotating, and there definitely is
(notwithstanding that they may be negligible in a given circumstance). If we choose to use or must use a
rotating frame, then the centrifugal and Coriolis forces are exact consequences of transforming the
equation of motion (Section 2.4). There is nothing ad hoc or discretionary about their appearance in the
rotating frame equation of motion, and there is no good in calling or thinking of the Coriolis force as a
’pseudo force’, or a ’fictitious correction force’ or any usage that might seem to question whether the
Coriolis force is a full-fledged member of the equation of motion.
2) The Coriolis and centrifugal forces preclude the conservation properties of an inertial
reference frame. Recall the elementary trajectory of Section 3.4; when observed from a rotating frame
the parcel veered to the right as expected from the deflecting Coriolis force. However, there was no
object in the environment that revealed a corresponding reaction force. Similarly, the parcel speed and
kinetic energy increased with time due to work by the centrifugal force, and yet there was no source for
the energy. The rotating frame equation of motion thus does not support global conservation of
momentum and neither does it preserve invariance to Galilean transformation. In practice these are not
serious failings. Of much greater importance is that the analysis and interpretation of geophysical flow
phenomena is far simpler when viewed from an Earth-attached, rotating reference frame. In the special
case of a reference frame attached to a rotating planet, the centrifugal term is cancelled by a small
deformation of the planet.
6 SUMMARY 56

3) The consequences of the Coriolis force include the possibility of inertial and geostrophic
motions. Because the atmosphere and the oceans are thin when viewed in the large and also stably
stratified, the horizontal component of winds and currents is generally much larger than the vertical
component. In place of the full three-dimensional Coriolis force, it is usually sufficient to consider only
the horizontal component of the Coriolis force acting upon the horizontal wind or currents, −f×V,
where f is the Coriolis parameter, f = 2Ωsin(latitude), times the local vertical unit vector. The Coriolis
parameter will have a very prominent part in the discussions that follow in Parts 2 and 3.
The analysis of a dense parcel released onto a slope (Section 5) revealed two kinds of motion that
depended directly upon the Coriolis force. There is a free oscillation, usually called an inertial
oscillation, in which an otherwise unforced current rotates at the inertial frequency, f . These inertial
oscillations are often a prominent feature of the upper ocean current following the passage of a moving
storm. A crucial, qualitative effect of rotation is that it makes possible a steady motion that is in balance
between an external force (wind stress or buoyancy force) and the Coriolis force. The characteristic of
this geostrophic motion is that the velocity is perpendicular to the applied force. It would be easy to
over-interpret results from a single parcel model, but a correct inference from this result is that Earth’s
rotation — by way of the Coriolis force — is indeed the basis for the slowly evolving winds and currents
that make up the majority of atmospheric and oceanic circulation systems outside of equatorial regions.

What’s next? This introduction to the Coriolis force continues (under a separate cover) with an
emphasis on the consequences for the atmosphere and ocean. Specific goals are to understand
What circumstances lead to a near geostrophic balance? in Part 2, and,
How do small but systematic departures from geostrophic balance lead to time-dependent, low
frequency motions, and east-west asymmetry? in Part 3.
The plan is to conduct a sequence of geostrophic adjustment experiments using a model of a single fluid
layer, often called the shallow water model. These experiments employ a sequence of approximations to
the latitudinal dependence of the Coriolis parameter, f , and are analyzed using potential vorticity
balance, among others. The tools and methods of Parts 2 and 3 are thus a considerable advance over
those employed here in Part 1, and are much more likely to be directly useful in your own research.
Everything that you have learned here in Part 1 regarding the Coriolis force acting on a single parcel is
applicable there as well. There is a lot that is new, including that in a fluid, an entire field of parcels will
often be moving in a way that is coupled (made coherent) by pressure gradients and hence is wave-like.
The properties of these waves depends very much upon our new/old friend, f .
7 SUPPLEMENTARY MATERIAL 57

7 Supplementary material

The most up-to-date version of this essay plus the related Matlab scripts may be downloaded from the
author’s public access web site: www.whoi.edu/jpweb/aCt.update.zip
Matlab scripts include the following:
rotation 1.m solves for the three-dimensional motion of a parcel as seen from an inertial and a rotating
reference frame. Used to make Fig. 8.
rotation 2.m solves for the two-dimensional motion of a parcel as seen from an inertial and a rotating
reference frame in which the centrifugal force is cancelled. Used to make Fig. 13. It is much easier to
see the results in the onscreen plotting vs. the hardcopy figure.
partslope.m solves for the motion of a single dense parcel on a slope and subject to buoyancy, bottom
friction and Coriolis forces as in Section 5. Easily configured to new experiments.

You might also like