Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Analytic formula for the geometric phase of an asymmetric top

Nicholas A. Mecholsky

Citation: American Journal of Physics 87, 245 (2019); doi: 10.1119/1.5093302


View online: https://doi.org/10.1119/1.5093302
View Table of Contents: https://aapt.scitation.org/toc/ajp/87/4
Published by the American Association of Physics Teachers
Analytic formula for the geometric phase of an asymmetric top
Nicholas A. Mecholskya)
Vitreous State Laboratory, The Catholic University of America, Washington, DC 20064
(Received 10 July 2018; accepted 11 February 2019)
The motion of a handle spinning in space has an odd behavior. It seems to unexpectedly flip back and
forth in a periodic manner as seen in a popular YouTube video (“Plasma Ben, Dancing T-handle in
zero-g, HD,” <https://www.youtube.com/watch?v¼1n-HMSCDYtM>). As an asymmetrical top, its
motion is completely described by the Euler equations and the equations of motion have been known
for more than a century. However, recent concepts of the geometric phase have allowed a new
perspective on this classical problem. Here, we explicitly use the equations of motion to find a closed
form expression for the total phase and hence the geometric phase of the force-free asymmetric top
and we explore some consequences of this formula with the particular example of the spinning
handle for demonstration purposes. As one of the simplest dynamical systems, the asymmetric top
should be a canonical example to explore the classical analog of the Berry phase. VC 2019 American
Association of Physics Teachers.
https://doi.org/10.1119/1.5093302

I. INTRODUCTION this can be viewed as an angular difference from the initial


state. This extra phase will be referred to as the “total phase”
The motion of spinning objects is ubiquitous. From a in this paper. The total phase is a combination of two parts.
seemingly mundane and classical motion, new mysteries are One is the dynamic part that depends on the time dynamics,
still being debated and surprising features are revealed.2–4 and the other is the geometric part, called the Hannay
One important observation is the identification of a Berry Angle,19 which is independent of time and hence “geometric.”
phase of rotating classical objects.5–7 Geometric phases have Alternatively, we may consider a periodic system whose
become popular recently due to their applications to material Hamiltonian is forced to change slowly due to some external
properties (see Ref. 8 for a recent review) and other theoreti- influence. An example of this could be the Foucault pendu-
cal and calculational advantages.9,10 Examples have relied lum.13,20 If the system is then returned to the original
on Berry’s original paper11 describing the quantum mechani- Hamiltonian, there is an extra phase due to this forcing. One
cal effect and relegating the classical version to the term might call these two examples the “passive” and “active”
“classical analog.” The classical analog of the Berry phase is cases.
the Hannay angle and was developed in the mid 1980s by In 1991, Richard Montgomery, and around the same time,
Hannay and others, inspired by Berry’s original work.6,7 The Mark Levi, published papers6,21 that derived a formula for
canonical example of the Hannay angle usually involves computing the total phase of rigid bodies
Foucault’s pendulum.12,13 However, even in the force-free
spinning of a top, the Hannay angle is apparent, yet not typi- 2ET
cally discussed in elementary treatments of the subject. Da ¼  X; (1)
M
Evidently, Jacobi was the first to write down the exact ana-
lytical expressions for the motion of the asymmetric top in where T is the period of the angular velocity vector in space
the 19th century. In Landau and Lifshitz’s Mechanics,14 it (here, we compute it from Eq. (16)) and X is the signed solid
was already appreciated that the motion of the top was not area swept out by the angular momentum vector. The
perfectly periodic in time. This was identified by Landau and dynamic part (2ET/M) is the integrated angle of the angular
Lifshitz when they stated that “this incommensurability has velocity projected onto the angular momentum vector. The
the result that the top does not at any time return to its origi- geometric part is a fraction of a unit sphere swept out by the
nal position” (see Ref. 14, Sec. 37, p. 120). However, this path of the angular momentum vector in the body frame
observation was not investigated further in that text and the (using the right hand rule to assign a sign).
connection to the Hannay angle was not included in their Even though much work has been done in this area in the
work which would have predated the original work by past few decades (see, for example, Ref. 22), the connection
Hannay7 and Berry.11 and explicit formula for the total phase of an asymmetric top
Geometric phases (Hannay angles) are not typically dis- has never been published. Part of the goal of this paper is to
cussed in undergraduate courses. The basic features and cal- take the last step and cast the exact expressions for the
culations can be seen in several references.5,7,15–18 The motion of the asymmetric top in the framework of the total
geometric phases of several real-world examples are calcu- phase. A closed form expression for the total phase will be
lated in Ref. 15. As a simple example, consider a classical presented and explored. We use the example of the T-shaped
system that has periodic motion. An example of this is the handle (seen flipping around on the International Space
asymmetric top in force-free motion in space. Some of the Station in a recent YouTube video1) to demonstrate some of
dynamic variables are periodic (like the angular momentum these expressions; however, the formulas are valid for a gen-
and angular velocity vectors). However, other variables (like eral asymmetric top. In some ways, this is the most elemen-
the Euler angles) are not necessarily periodic with the same tary example of the total phase available in classical physics.
period. In one cycle of one set of periodic variables, the sys- Even in this simple system, a wide variety of questions and
tem almost returns to its original state except for the varia- lines of investigation are possible. To that end, an explora-
bles that are not periodic with the same period. Typically, tion and some observations of the formula are in order.

245 Am. J. Phys. 87 (4), April 2019 http://aapt.org/ajp C 2019 American Association of Physics Teachers
V 245
This paper begins by reviewing the dynamics of the asym-
metric top. The angular velocity and Euler angles are deter-
mined as functions of time. From these expressions, it is
found that the total phase is determined directly and com-
pared with Montgomery’s formula. Finally, some observa-
tions about the total phase are made and some conclusions
are discussed.

II. DYNAMICS AND TOTAL PHASE


As an example for visualization, and without loss of gen-
erality, we may use a T-shaped handle to explore our dynam-
ics such as the dynamics seen in the video.1 For any three Fig. 1. Two equivalent rigid bodies of equal uniform density. A T-shaped
handle (green) with l1 ¼ 8, l2 ¼ 4, and w ¼ 1 and the equivalent solid ellip-
physical moments of inertia, there is a corresponding T- soid that has the identical principal moments of inertia given by Eq. (3),
shaped handle, and thus, we are losing no generality. with semi-axes (a, b, c) ¼ (4.23281, 3.07318, 0.645497). The ellipsoid is
Consider a handle of uniform density and unit mass with a shrunk by 25% for presentation.
cross piece of length l1, a shaft of length l2, and a square
cross-sectional width of w (Fig. 1). from the eigenvalues) has unit magnitude. The moment of
The total moments of inertia can be computed using the inertia for this particular choice is given by
parallel axis theorem with individual square prisms. This
will give a moment of inertia tensor with diagonal elements I1 ¼ b2 þ c2 ; (4a)
6l1 l2 w 2
þ 2l21 w2 þ l22 wð6l1 þ wÞ þ l42 þ 4l1 l32 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
I1 ¼ 2
; (2a) I2 ¼ b2 þ 2  3b4  2b2 c2  3c4 þ c2 ; (4b)
12ðl1 þ l2 Þ 2
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
ðl1 þ 2l2 Þw2 þ l31 I3 ¼ b2 þ 2  3b4  2b2 c2  3c4  c2 ; (4c)
I2 ¼ ; (2b) 2
12ðl1 þ l2 Þ
    with the third semi-axis of the ellipsoid given by
l1 l2 l21 þ 5w2 þ 4l22 þ l21 l21 þ w2 þ l22 wð6l1 þ wÞ þ l42
I3 ¼ ; pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
12ðl1 þ l2 Þ2 2  3b4  2b2 c2  3c4  b2  c2
a¼ pffiffiffi : (5)
(2c) 2
where the coordinate axes have been chosen to have a diagonal This particular choice of semi-axes of the ellipsoid is always
moment of inertia tensor. An ellipsoid (with semi-axes lengths chosen to be labeled so that I1 < I2 < I3. However, as long as
a, b, and c) with the same moment of inertia is given by a > b > c in the ellipsoid, this will always be the case.
rffiffiffi In Sec. II A, we take a small digression to identify the full
5pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi space of moments of inertia.
l2 w2 þ l3
3 pffiffiffiffiffiffiffiffiffiffiffiffiffi 1
a¼ ; (3a)
2 l1 þ l2 A. Space of possible moments of inertia
qffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  ffi
5 2 þ 2w2 þ l2 w2 þ l4 Not all triplets (I1, I2, I3) of positive numbers are valid
3 2l 1 l2 3l 2 w þ 2l 2 1 2 moments of inertia. To have a valid (physical) moment of
b¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; (3b)
inertia, we must satisfy the following relations:
2 ðl1 þ l2 Þ2
rffiffiffi I1 þ I2  I3 ; (6)
1 5
c¼ w; (3c) I2 þ I3  I1 ; and (7)
2 3
I3 þ I1  I2 ; (8)
shown superimposed in Fig. 1. Although this is correct for
all l1, l2, and w, here we wish l1 to be large enough (other-
which may be referred to as the inertial inequalities.
wise, a relabeling of a, b, and c is needed so that the Without loss of generality (we may relabel axes in the
moments of inertia are correctly ordered) with respect to l2 object if necessary), we may further restrict ourselves to the
(and w small enough) so that a > b > c. region, I1 < I2 < I3. For labeling purposes, consider only the
It will turn out that the only important parameters for the
moments of inertia where I12 þ I22 þ I32 ¼ 1. In this case (we
description of the body are those that determine the direction
will see that these are the only cases we need to consider),
that the moment of inertia points in the moment of inertia
we can show the region of possible physical moments of
space (the polar and azimuthal angles relative to the I1, I2,
inertia as those moments of inertia inside the red boundary in
and I3 axes). Thus, the absolute magnitude of the vector of
the I1–I2 plane in Fig. 2. The curves that bound this region
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

the eigenvalues of the moment of inertia is not important.
We will find it useful to use a moment of inertia that is are the curve of all prolate spheroids ðI2 ¼ ð1  I12 Þ=2Þ,
parameterized by two variables, b and c, the semi-axes of a the curve of all oblate spheroids (I2 ¼ I1), p and a curve
ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi of
solid ellipsoid. The third semi-axis of this ellipsoid is deter- degenerate ellipsoids where c ¼ 0, I2 ¼ 12 ð 2  3I12  I1 Þ.
mined such that the moment of inertia vector (composed Thus, picking 0 < c < b < 1 fixes a particular asymmetric

246 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 246
angular momentum vector executes periodic motion, where
the origin is fixed and the terminus sweeps out a curve in the
body frame. The curve that the angular momentum sweeps
out in space is constrained to be the intersection of these two
constants of motion. This intersection curve is called the pol-
hode. If we use Mi ¼ Iix0i, then we may rewrite Eqs. (10)
and (11) slightly to give

M2 ¼ M12 þ M22 þ M32 ; (12)


 
1 M12 M22 M32
E¼ þ þ ; (13)
2 I1 I2 I3

which are the equivalent to Eqs. (5.45) and (5.46), pp.


203–204 of Ref. 23, or Eqs. (37.3) and (37.4), p. 116, and
Sec. 37 of Ref. 14. This intersection can be seen in Fig. 3(a).
Equations (12) and (13) represent a sphere of constant
magnitude (the angular momentum magnitude is fixed) and
an intersecting ellipsoid that is the surface of constant energy
in angular momentum coordinates (a space where the coordi-
nates are the projections of the angular momentum along the
Fig. 2. Full parameter space of asymmetric rigid bodies of fixed moment of body’s principal moments of inertia) called the Binet
inertia magnitude equal to 1. I1 and I2 are two of the principal moments of iner-
tia labeled so that I1 < I2 < I3, and the magnitude of the moment of inertia vec-
ellipsoid.23
tor is unity. The red curves are degenerate asymmetric rigid bodies, here The two constant surfaces must intersect for a physical
modeled as solid ellipsoids: Prolate spheroids (rod-like shapes), spheres, oblate object, and this means that the constant angular momentum
spheroids (coin-like shapes), and degenerate ellipsoids (lines and disks). The sphere is smaller than the largest Binet ellipsoid axis and
equations for these curves are given in the text. The contours are curves of con- bigger than the smallest Binet ellipsoid axis. This gives the
stant minimum values for the total phase. Given a body (parametrized by two condition
of the three values of the moment of inertia), any initial condition will lead to at
least a total phase of the given value indicated by the contour values.
2EI1  M2  2EI3 : (14)
top. For the rest of this paper, we assume a rigid object with
moments of inertia fixed. This will always occur since the alternative is for no inter-
section of the two constant surfaces and thus no physical
B. Exact dynamics of the asymmetric top dynamics. There are two possible types of curves on the
angular momentum sphere and the Binet ellipsoid (not
Suppose we decide to spin this asymmetric top around
including degenerate cases). To avoid degenerate cases (dis-
some axis with some initial angular velocity and without any
cussed elsewhere, for example, in Ref. 14 Sec. 37, p. 119, or
external torques. Let us say that in the body frame of refer-
Ref. 23), we shall always assume strict inequalities. We shall
ence (determined by the principal moments of inertia
also take M2 > 2EI2 for the equations below, but for the case
I1 < I2 < I3 in a right handed coordinate system), the initial M2 < 2EI2, the formulas require 1 $ 3 in the indices.
angular velocity is given by The dynamics of the angular velocity in the body frame
x0 ¼ ðx01 ; x02 ; x03 Þ: (9) evolve according to the Euler equations

I1 x_ 1 ðtÞ ¼ ðI2  I3 Þx2 ðtÞx3 ðtÞ; (15a)


We will show below that the total phase will not depend on
the magnitude of x0 . Thus, we shall take jx0 j ¼ 1. I2 x_ 2 ðtÞ ¼ ðI3  I1 Þx3 ðtÞx1 ðtÞ; (15b)
The time dynamics of the angular velocity is then deter-
mined by these initial conditions. In torque-free motion, there I3 x_ 3 ðtÞ ¼ ðI1  I2 Þx1 ðtÞx2 ðtÞ: (15c)
are two constants of motion for the asymmetric top. One is the
total angular momentum of the top, which is given by This may be solved exactly in terms of Jacobian elliptic
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi functions (Eq. (37.10) of Ref. 14, Sec. 37, p. 118). The
motion of the angular velocity vector is perfectly periodic in
M ¼ jI  x0 j ¼ I12 x201 þ I22 x202 þ I32 x203 ; (10)
time with period
 
and the other is the total energy of rotation (assume that the
4 K kðE; M; I1 ; I2 ; I3 Þ2
center of mass velocity of the asymmetric top is 0), which is T ðE; M; I1 ; I2 ; I3 Þ ¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with
given by ðI3  I2 ÞðM2  2EI1 Þ
1 1  I1 I2 I3
E ¼ x0  I  x0 ¼ I1 x201 þ I2 x202 þ I3 x203 : (11)
2 2 (16a)
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
The motion of the angular momentum vector in space coor- ðI2  I1 Þð2EI3  M2 Þ
dinates is, of course, fixed since no torque is applied to kðE; M; I1 ; I2 ; I3 Þ ¼ ; (16b)
change the angular momentum. In the body coordinates, the ðI3  I2 ÞðM2  2EI1 Þ

247 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 247
convention, two of the Euler angles (h and w) are given
algebraically in terms of the angular velocity functions, and
the third (/) is given as a first-order differential equation in
time. Reference 14 claimed that this may be further inte-
grated to give / in terms of theta functions, referencing
Chapter 6 of Ref. 24 and arriving at Eqs. (37.18)–(37.20).
However, we were not able to verify this solution due to
differences in the theta function notation in many sources.
However, a straightforward solution is provided in the
Appendix for all Euler angles in Eqs. (A4) and (A12).
Briefly, the integration for / may be carried out directly to
give Incomplete Elliptic Integrals of the Third Kind. See
the Appendix for details.

C. Formula for the total phase


Even though the equation for the angular velocity is
exactly periodic in time, the orientation and angles after a
period are not periodic and will not in general return to the
same orientation. This angular mismatch after a period of the
angular velocity will be called the total phase (or the same
quantity mod 2p). The term “geometric phase” (or the
Fig. 3. (a) Intersection in angular momentum space of two constants of the
motion: the constant moment of inertia sphere (blue) and the energy ellip-
Hannay angle) is used to refer to just the geometric part of
soid (yellow). (b) Representative motion of the body angular velocity in this phase.6,19,25 We reserve the term “Berry phase” as the
time. The solid red (dashed green, dot-dashed blue) curve plots the value of quantum mechanical version of the geometric phase where a
the angular velocity component along the body x (y, z) axis which corre- wave function is explicitly transported in parameter space. In
sponds to the smallest (intermediate, largest) moment of inertia. We choose contrast, we investigate the natural (so to speak) phase asso-
the total mass (from a uniform density) of the T-handle in Fig. 1 to be such ciated with force-free evolution.
that the moment of inertia has unit magnitude. The initial angular velocity
has unit magnitude, and the direction is given by a polar angle (measured
In computing the total phase, we could use Montgomery’s
from the body z axis), hx ¼ p/2–0.15, and an azimuthal angle, /x ¼ p/2–0.2, formula6 (Eq. (1))
measured from the body x axis. This sets the moment of inertia and total
rotational energy. Those, along with the moments of inertia, are E 2ET
¼ 0:264805 J; M ¼ 0:533252 Js; I1 ¼ 0:286827 Js2 ; I2 ¼ 0:533256 Js2 , and Da ¼  X; (18)
I3 ¼ 0:795844 Js2 . The period of the motion is 23.8181 s, marked with a M
gray vertical line in (b).
where T is the period from Eq. (16) and X is the signed solid
where K(m) is the Complete Elliptic Integral of the First area swept out by the angular momentum vector. The
Kind, namely, dynamic part ð2ET=MÞ is the integrated angle of the angular
velocity projected onto the angular momentum vector over
ð p=2 one period. The geometric part is the fraction of a unit sphere
dh
K ðm Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (17) swept out by the angular momentum vector over one period
0 1  m sin2 h (using the right hand rule to assign a sign). This might be a
difficult computation to do since it is a geometric surface
In some notations, the argumentpffiffiffi offfi the Complete Elliptic area enclosed by the polhode. It may still be done
Integral of the First Kind, K, is m and not m. We choose numerically.
the notation in Eq. (17), but care must be taken here. Note However, considering that we have exact expressions for
that when M2 < 2EI2, the replacement 1 $ 3 must be made the Euler angles (Eqs. (A4) and (A12)), we may evaluate the
in the above quantities. This has the effect of making T
total phase directly. The angular momentum vector after one
always real and 0 < k < 1.
period will be identical to its starting value.6 Thus, the angu-
Given the initial angular velocity vector x0 in the body
lar difference around this axis is the total phase. Consider
frame with polar angle hx from the body z axis and azi-
that in some cases (when M2 > 2EI2), this may involve w as
muthal angle /x from the body x axis, the initial angular
well as /, but not h. In this case, w always changes by 2p.
momentum (M0 ¼ I  x0 ) is not collinear with the angular
When M2 < 2EI2, w is perfectly periodic, and thus, the angu-
velocity (since I1 6¼ I2 6¼ I3). The initial angular momentum
lar change in / is zero. If we assume that the Euler angle /
is fixed in the space frame, and we may define the z axis
at time t ¼ 0 is 0, then we have the change in angle, Da, over
of the space coordinates to be this angular momentum
one period which is
vector.
Since the angular velocity vector is explicitly determined (
in time (See Appendix Eqs. (A1) and Ref. 14 Eq. (37.10)), /ðTÞ  2p M2 > 2EI2
we may further express the Euler angles as explicit func- Da ¼ (19)
/ðTÞ M2 < 2EI2 :
tions of time. Here, we choose the Euler angle convention
of both Refs. 23 and 14 (Figure 4.7, p. 152 and Fig. 47 Sec.
35, respectively). This describes a sequence of intrinsic ele- Substituting the expressions for / in the Appendix (Eqs.
mental rotations about the body z–x0 –z00 axes. In this (A12)), Eq. (19) reduces to

248 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 248
8  
>
> I3 ðI2  I1 Þ 2
>
> P  jk
>
> MT 4MðI3  I1 Þ I1 ðI3  I2 Þ
>
> þ s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  2p M2 > 2EI2
>
> I3 I1 I3
>
> ð I  I Þ ð M 2
 2EI Þ
>
>
3 2 1
>
< I1 I2 I3
Da ¼ ! (20)
>
> I3 ðM2  2EI1 Þ 2
>
> P  jk
>
> MT 4M ð I  I Þ I1 ð2EI3  M2 Þ
>
> þ
3 1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi M2 < 2EI2 ;
>
>
>
> I3 I1 I3 ðI2  I1 Þð2EI3  M2 Þ
>
>
>
: I1 I2 I3

where PðnjmÞ is the Complete Elliptic Integral of the Third In Sec. II D, to check Eq. (20), we evaluate it by calculat-
Kind defined by ing Eq. (1) numerically and by solving for a rotation matrix
ð p=2 after one period.
dh
PðnjmÞ ¼ ; (21)
0 ð1  n sin hÞð1  m sin2 hÞ1=2
2
D. Demonstration of the formula
where k is defined in Eq. (16b).
For illustrative purposes, we choose a handle shape where
It is worthwhile to note that the total phase formula in Eq.
l1 ¼ 8, l2 ¼ 4, and w ¼ 1 as shown in Fig. 1. We further choose
(20) seems to depend on E, M, I1, I2, and I3. However, it can
the total mass (from a uniform density) to be such that the
be shown that it does not depend on the magnitude of the
moment of inertia has unit magnitude. The initial angular
moment of inertia (I 2 ¼ I12 þ I22 þ I32 ) nor on the magnitude
of the initial angular velocity. velocity is chosen so that it has unit magnitude, and the direc-
To see the independence of magnitude I, note that the tion is given by a polar angle (measured from the body z axis),
energy, Eq. (11), is homogeneous of order 1 in I. The angular hx ¼ p=2  0:15, and an azimuthal angle, /x ¼ p=2  0:2,
momentum, Eq. (10), is also homogeneous of order 1 in I. measured from the body x axis. This corresponds to a large ini-
The functions k and T of Eq. (16) are therefore homogeneous tial velocity near the middle moment of inertia, I2. This
of order 0 in I. It then follows that Eq. (20) is independent of sets the total angular momentum and total rotational energy.
the magnitude of the moment of inertia. Those, along with the moments of inertia, are E ¼ 0:264805 J;
For the independence of the magnitude of x0 , let us con- M ¼ 0:533252 Js; I1 ¼ 0:286827 Js2 ; I2 ¼ 0:533256 Js2 , and
sider the initial polar and azimuthal angles of the angular
velocity (x0 ) with respect to the body axes (hx and /x )
fixed for the moment. Note that in this case, the angular
momentum is homogeneous of order 1 in jx0 j, and the
energy is homogeneous of order 2 in jx0 j. This means
that k is homogeneous of order 0 in jx0 j and T is homoge-
neous of order 1 in jx0 j. Then, looking at the total
phase formula, Eq. (20), this means that the total phase is
independent of the magnitude of the initial angular veloc-
ity, jx0 j.
Considering that the total phase formula is itself a con-
stant of motion (it is composed of constants of motion),
the initial angles, hx and /x , which set the constants of
the motion (E and M) along with the direction of the
moment of inertia (or simply I1 and I2 along with the con-
straint that I12 þ I22 þ I32 ¼ 1), are the only parameters
needed to determine the total phase. Thus, every point
along the closed curve on the Binet ellipsoid has the same
total phase since every point has the same quantities of E,
M, I1, I2, and I3.
Thus, curves of the same total phase foliate the Binet ellip-
soid. An example (we use Js2 in the units of the moment of Fig. 4. Total phase (mod 2p) of the body given in Fig. 1 for an asymmetric
inertia as opposed to kg m2 to make transparent the relation- top with (I1, I2, I3) ¼ (0:286827 Js2 ; 0:533256 Js2 ; 0:795844 Js2 ). For each
ship between the energy (J), angular momentum (Js), initial angle, a given curve on the Binet ellipsoid is represented. The Binet
moments of inertia (Js2), and angular velocity (s1) in Eqs. ellipsoid is the ellipsoid of constant energy in the angular momentum space.
Given an initial angular momentum, the resulting dynamics traces out a
(16) and (20)) of this can be seen in Fig. 4 for the case of (I1, closed curve on this surface. The final angular mismatch (mod 2p) is the
I2, I3) ¼ (0.286827 Js2, 0.533256 Js2, 0.795844 Js2). Here, total phase, and the corresponding colors are related to the corresponding
the color of the curve corresponds to the total phase modulo values from 0 to 2p. Note that there are multiple curves where the phase is
2p. near 0.

249 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 249
I3 ¼ 0.795844 Js2. The dynamics of x in the body frame are Binet ellipsoid (Fig. 4). Without loss of generality, we
shown in Fig. 3(b). In this case, the period calculated from may choose the initial conditions to lie on the body x–z
Eq. (16a) is 23.8181 s. We can see from the shape of the plane so that the azimuthal angle of the initial angular veloc-
curves that this corresponds visually to the period of Fig. ity /x is equal to 0.36 Thus, for the remainder of this paper,
3(b). One aspect to note from this plot is that the dashed we take
green curve shows that the body y-component of the angu-
lar velocity (the axis of the intermediate moment of inertia) x0 ¼ ðsin hx ; 0; cos hx Þ: (23)
suddenly reverses twice describing the flip known as the
Dzhanibekov effect or the tennis racket theorem.14,23,26,27 If we consider all combinations of moments of inertia in
Interestingly, most classic text in dynamics describes this the region indicated in Fig. 2, along with all initial
behavior as “unstable;” however, the motion is completely 0 < hx < p/2, this covers all initial conditions (using symme-
determined and is periodic, eventually coming back to the try of the angle, direction of spin, and independence of jx0 j
unstable point. and jIj). Equation (20) has been checked exhaustively
We may now compute some of the quantities that have against Eq. (1) for 292 sets of moments of inertia equally
been developed. The dynamic part of the total phase (2ET/ spaced throughout the region in Fig. 2 with 22 angles
M) from Eq. (1) is 23.6555. If we integrate the projected between 0 < hx < p=2 for each moment of inertia choice.
angular velocity (from Eq. (A1) or numerically integrate Eq. This corresponds to 6446 initial conditions. In all cases, both
(15)) along the angular momentum axis for one period, we Eqs. (20) and (1) give the same value.
get
ðT
xM III. DISCUSSION OF THE FORMULA
dt ¼ 23:6555; (22)
0 jMj A. Symmetries of the total phase
which matches the dynamic part. If we compute the solid As mentioned before, the total phase formula has notable
angle of intersection between the sphere of angular symmetries independence from jIj
including itspffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
momentum and the Binet ellipsoid numerically using ¼ I12 þ I22 þ I32 and jx0 j ¼ x201 þ x202 þ x203 .
Mathematica (seen in Fig. 3(a)), we get the signed area Independence from jIj follows from the observation
X ¼ 2.1253. that jIj is directly proportional to the total mass of a rigid
This gives a calculated total phase of 21.5302. The for- body. Changing jIj amounts to changing the mass scale. If
mula from Eq. (20) gives a value of 21.5303. The difference the mass scale were changed, the dynamics would
is mostly due to the numerical discretization of the area of be unchanged, especially the geometric quality of the
intersection and can be made to be smaller with a finer rotation.
discretization. To qualitatively understand the independence of Eq. (20)
Another way of checking both formulas is by numeri- from jx0 j, we see that changing the magnitude of jx0 j
cally integrating the Euler equations over one period and changes the timescale for the dynamics but, geometrically,
solving a system of equations for the rotation matrix after does not affect the phase. If we had made a video recording
one period. The rotation matrix would give the geometric of the spinning and played it back at a slower speed, the
rotation of the initial system compared to the final orienta- geometrical quality of the rotation would also remain
tion after one period. In this case, the rotation around the unchanged.
angular momentum vector is given by 2.68074 which is Thus, we only need three parameters to specify the total
precisely the value of the total phase formula (21.5303), phase: two moments of inertia, I1 and I2 (or two ellipsoidal
modulo 2p. semi-axes (b and c)), and the polar angle of the initial angu-
To further simplify the total phase formula, we may use lar velocity in the body frame, hx .
the observation that there are related initial conditions that We may explicitly substitute the expressions for E and M
describe the same total phase (since the total energy and (Eqs. (11) and (10)) along with Eq. (23) and the definitions
magnitude of the angular momentum are the same through- for T and k (Eq. (16)) into the total phase formula (Eq. (20))
out the evolution). These lie on the same closed curve on the to give

8 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  !!
>
> I2 I32  I12 ð1  sec2 hx Þ   I 2
>
> 4 b 2 2 3 2
I1 K b tan hx þ ðI3  I1 ÞP  2 b b tan hx  2p M2 > 2EI2
>
>
< I1 I3 ðI3  I1 ÞðI2  I1 Þ I1
Da ¼
> 2  ffi0
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ! !1 (24)
>
>
2
I2 I1 þ I3 cot hx @ 2 2
cot hx 2 2 2
I cot hx cot hx A
>4
> I1 K þ ðI3  I1 ÞP  3 2 b M2 < 2EI2 ;
>
: I12 I3 ðI3  I1 ÞðI2  I1 Þ b I1 b

where

I 1 ðI 2  I 1 Þ
b¼ ; (25)
I 3 ðI 3  I 1 Þ

250 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 250
and K and P are defined in Eqs. (17) and (21), respectively.
we mightffi have eliminated I3 as well using
Here,pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
I3 ¼ 1  I12  I22 , but that would perhaps not benefit the
exposition.

B. Minimum total phase


Figure 5 shows two plots for the above demonstrated case.
The solid red curve is the total phase for a given initial con-
dition described by the polar angle (hx ) of the angular veloc-
ity with respect to the body z axis and an azimuthal angle of
zero (/x ) with respect to the body x axis. The total phase
(solid red) is the difference between the dynamic phase (dot-
ted green) and the geometric phase (dot-dashed blue). The
dashed black curve is the corresponding time period, T, for
the angular velocity or angular momentum vectors. For a Fig. 6. Total phase for moments of inertia given by (I1, I2, I3) ¼ (0.26 Js2,
0.55 Js2, 0.793662 Js2). The initial angular velocity angle is given as a polar
given object, the total phase, dynamic phase, and period
angle from the body z axis. When the total phase is an integer multiple of
diverge at an angle that corresponds to the separatrix for the 2p, the herpolhode (curve of the terminus of the angular velocity in space
intersection of the Binet ellipsoid and the moment of inertia coordinates) is a close plane curve. Various examples are plotted near the
sphere. The critical angle is given by the condition intersection of multiples of 2p and the total phase curve.

M2 ¼ 2EI2 ; (26) M2 ¼ 2EI2, the total phase increases without bound. One
question that is apparent is what is the minimum total phase
which reduces to for a given object? For example, are there objects (I1 and I2
pffiffiffi or b and c values) where there are no initial conditions that
hcrit ¼ cot1 ð bÞ; (27) give rise to a total phase of exactly 2p?
In Fig. 2, we see that the minimum total phase can be as
with b defined in Eq. (25). high as desired. Looking at all objects, the largest minimum
This critical angle is displayed as a gray vertical line in total phase is unbound for objects with moments of inertia
pffiffiffi
Figs. 5 and 6. Note that the minimum value for the total approaching I1 ¼ I2 ¼ I3 ¼ 3=3. There is an apparent par-
phase occurs at the extremes of 0 and p. This appears to be a adox in this behavior. If all moments of inertia are equal and
generic behavior in all the initial conditions investigated. It the total angular momentum is fixed, then the body spins
might not be too difficult to differentiate Eq. (24) and inves- around whatever axis it is initially spinning about. The total
tigate the sign of the derivative, but may not add much. phase would seem to be zero. However, the total phase for-
Additionally, near the critical angle, the total phase diverges. mula goes to infinity. The resolution may be seen in the
When the total phase modulo 2p is equal to 0, trajectories behavior ofpthe ffiffiffi period T. When each moment of inertia
that are started on these curves are closed and each Euler approaches 3=3, the period is seen to diverge. Thus, with T
angle (mod 2p) is periodic with a period whose integer mul- going to infinity, the total phase is permitted to pffifficorrespond-

tiple is the period of the angular velocity vector. As the tra- ingly diverge. Note that when I1 ¼ I2 ¼ I3 ¼ 3=3 exactly,
jectories on the Binet ellipsoid approach the curve the object is no longer dynamically asymmetric (an example
of this is a sphere).
On the other hand, the asymmetric top with the smallest
minimum possible total phase for any initial condition is a
total phase of 0 for objects with pffiffiffi moments of inertia
approaching I1 ¼ 0 and I2 ¼ I3 ¼ 2=2. An example of this
is a thin solid rod.
If we look at initial conditions near the body z or x axes
(hx  0 or hx  p=2, respectively), we may work out the
minimum total phase directly from linearizing the Euler
equations (Eq. (15)). The Euler equations reduce to a simple
harmonic oscillator equation where the period may be identi-
fied directly. Since in this case, the geometric part of the total
phase is zero (the angular momentum does not enclose any
solid angle), the total phase is given by the dynamic part, for
which there is a simple formula (see Eq. (1)).
In the case where hx  0, the minimum total phase
Fig. 5. For the demonstrated example of Fig. 3, the solid red curve indicates
reduces to
the total phase for different initial conditions with varying polar angles of sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the angular velocity. A polar angle of 0 is an angular velocity along the I1 I2
body z axis. The total phase (solid red) is the difference between the
Damin;z ¼ 2p : (28)
ðI3  I2 ÞðI3  I1 Þ
dynamic phase (dotted green) and the geometric phase (dot-dashed blue).
Here, we see that regardless of the initial condition, the total phase will be
larger than about 2p radians. The black dashed curve plots the corresponding In the case where hx  p=2, the minimum total phase is
period of the angular velocity or angular momentum vector in seconds. given by

251 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 251
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
I3 I2 Large nuclei such as 135 Pr and 163 Lu provide examples of
Damin;x ¼ 2p : (29) asymmetric top systems in nuclear physics.32,33 Through X-ray
ðI 2  I 1 ÞðI 3  I 1 Þ spectroscopy, the so-called “wobble” bands provide a way to
measure angular momentum transitions of spinning nuclei. The
These formulas may also be derived by expanding Eq. (24) connection to closed total phase orbits still needs to be made.
near hx  0 and hx  p=2. It is worth noting that the total
phase near these extremes is composed of entirely the V. CONCLUSIONS
dynamic part of the total phase. For initial conditions away
from these extremes, the total phase is a combination of the In this paper, we have found the total phase formula (and
dynamic and geometric parts as can be seen in Fig. 5. thus trivially the geometric phase) explicitly in terms of special
The locus of points where Eqs. (28) and (29) are equal is functions. This leads immediately to observations about the
exactly the location of the kinks in the contours of Fig. 2. total phase for the asymmetric top in force-free motion. The
total phase of a rigid object is independent of the magnitude of
C. Closed herpolhodes the moment of inertia, independent of the magnitude of the
angular velocity, and all initial conditions along the path of the
We may also use the total phase formula (Eq. (20)) to pro- angular velocity vector in space have the same total phase and
duce curves of closed herpolhodes. A herpolhode is the curve geometric phase. Thus, for a given rigid body, the total phase,
traced out by the terminus of the angular velocity vector in dynamic phase, and geometric phase only depend on a single
the space coordinates. In general, these do not close after a angle, the polar angle that the initial angular velocity vector
period. However, if we use initial conditions that have a total makes when it crosses the body x–z plane (for instance).
phase that is an integer multiple of 2p, we find interesting Additionally, in the parameter space of the moments of inertia,
closed curve patterns. each class of asymmetric rigid bodies has a minimum total
Figure 6 shows the total phase for all initial polar angles phase, that is, a minimum angle by which an object must rotate
(hx ) of the angular velocity (with /x ¼ 0). The particular after undergoing one period of the angular velocity vector in
object (whose moments of inertia are given by (I1, I2, space. For some objects, the total phase can be as high as
I3) ¼ (0.26 Js2, 0.55 Js2, 0.793662 Js2)) has initial conditions desired. Closed trajectories were also highlighted, and closed
with closed herpolhode trajectories. The corresponding herpolhodes were illustrated and discussed briefly.
angles that result in the smallest closed herpolhodes are
(from left to right) 0.9562, 1.004, 1.011, 1.012, 1.014, 1.021, ACKNOWLEDGMENTS
1.061, and 1.2742 radians. These are exactly the angles
where the total phase is a multiple of 2p. In this case, the The author would like to thank Richard Cecil for many
multiples are 2, 3, 4, 5, 5, 4, 3, and 2, respectively. Note that conversations related to the geometric phase and other
closed herpolhodes are possible for all integer multiples of topics. Additionally, the author would like to thank Richard
2p greater than 2. The critical angle is 1.01262. Some of the Montgomery for many observations and suggesting the line
closed herpolhode figures are depicted in Fig. 6. of analysis for Eqs. (28) and (29) during the review of this
Two other comments about closed herpolhodes should be manuscript. This work was funded by the Vitreous State
made. There is vertical symmetry for all herpolhodes in the Laboratory.
plot, but for the even multiples of 2p, (4p, 8p, 12p, etc.), the
herpolhodes also have a horizontal symmetry. Additionally, APPENDIX: EXPLICIT DERIVATION OF THE TIME
changing the magnitude of jx0 j expands or contracts the her- DEPENDENCE OF THE EULER ANGLES
polhode image; but otherwise, it remains unchanged.
It is unclear what, if any, connection these closed curves The solution of the time dependence of the angular veloc-
and the associated energies have to the Bohr-Sommerfeld ity for an asymmetric top in force-free motion is given in Eq.
quantization in quantum mechanics. (37.10) of Ref. 14, reproduced here for convenience
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
IV. APPLICATIONS AND POSSIBLE CONNECTIONS 2EI3  M2  
x1 ðtÞ ¼ cn Dtjk2 ; (A1a)
I 1 ðI 3  I 1 Þ
The primary purpose of this paper was to use the asym-
metric top as an intuitive and natural example of the total sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
phase. However, we give three examples of the connections 2EI3  M2  
x2 ðtÞ ¼ sn Dtjk2 ; (A1b)
that could be made with other areas of physics. I 2 ðI 3  I 2 Þ
Quantization of the asymmetric top28 leads to quantized sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
energy levels of the quantum mechanical asymmetric top. M2  2EI1  
How this corresponds to the closed orbits of the last section x3 ðtÞ ¼ dn Dtjk2 : (A1c)
I 3 ðI 3  I 1 Þ
still needs to be studied.
When molecules are no longer symmetric, the asymmetry
Here, snðujk2 Þ is the Jacobi Elliptic Sine function defined by
leads to observable spectroscopic modes that allow for anal-
inverting the Jacobi Elliptic Integral of the First Kind, F,
ysis and characterization29–31 (see Chap. 1, Sec. 4 of Ref. 30
through the pair of equations
and Chap. 4 of Ref. 31). Understanding the dynamics of the
asymmetric top classically gives us an understanding of the
connection with the rotational modes of an asymmetric mol- sin / ¼ snðujk2 Þ; (A2a)
 2 ð/
ecule. Making further connections with the total phase aris-
ing from simple rotation of these molecules could lead to dt
u ¼ F /jk ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; (A2b)
further insight. 0 1  k2 sin2 t

252 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 252
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cnðujk2 Þ ¼ 1  sn2 ðujk2 Þ, and dnðujk2 Þ ¼ 1  k2 sn2 ðujk2 Þ. When we integrate this, we get
The definition of D is given in Eq. (A6d), and k is defined in ðt   ð Dt
Eq. 16(b) of the main text. d/ M A du
The above equations (Eq. (A1)) are valid when M2 > 2EI2. /¼ dt ¼ t þ 2 sn2 ðujj2 Þ
;
0 dt I 3 BD 0 1  a
When M2 < 2EI2, all indices 1 $ 3 (including in the definition
(A7)
of k and x). Alternatively, we may use Eq. (A1) for all initial
conditions without making any index changes (extending with
k > 1 in the region M2 < 2EI2). This is the most convenient,
but care must be taken. For the rest of the Appendix, we use C
this extended definition of k to simplify the notation but at the a2 ¼  : (A8)
B
end of the Appendix produce equations that are valid in the
notation of the text. To distinguish it, we use j. Thus We may use Eq. (400.01) (in the definition of the Incomplete
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Elliptic Integral of the Third Kind of Ref. 34)
ðI2  I1 Þð2EI3  M2 Þ

ðI3  I2 ÞðM2  2EI1 Þ
; (A3)  2 2
 ðu dx
P a ; /jj ¼ 2 sn2 ð xjj2 Þ
; (A9)
0 1  a
for all I1 < I2 < I3 and any E and M.
We may further express the Euler angles as explicit func- where u is defined in Eq. (A2a), to do the above integral.
tions of time. We choose the Euler angle convention of both The exact expressions for each region of initial conditions
Refs. 23 and 14 (Fig. 4.7, p. 152 and Fig. 47 Sec. 35, respec- (M2 > 2EI2 and M2 < 2EI2) must be done with care to avoid
tively). This describes a sequence of intrinsic elemental rota- problems with the definitions of built-in functions.
tions about the body z  x0  z00 axes. In this convention, two For the case where initial conditions give M2 > 2EI2
of the Euler angles (h and w) are given in terms of the angu- (0 < j < 1), we get
lar velocity functions (see Ref. 14, Eq. (37.14))
M M ðI 3  I 1 Þ
I3 x 3 /ðtÞ ¼ tþ P
cos h ¼ ; (A4a) I3 I1 I3 D
M  
I 3 ðI 2  I 1 Þ  
I1 x 1   ; am Dtjj2 jj2 ; (A10)
tan w ¼ : (A4b) I 1 ðI 3  I 2 Þ
I2 x 2
where amðujmÞ is the Jacobi Amplitude. If u ¼ Fð/jmÞ, then
The final Euler angle, /, is given as a first-order differen- / ¼ amðujmÞ, where F is the Incomplete Elliptic Integral of
tial equation as follows. Using Eq. (37.16) of Ref. 14, we the First Kind defined in Eq. (A2b). Again, j is defined in
may substitute in the expressions of Eq. (37.6) of Ref. 14. Eq. (A3), and D is given in Eq. (A6d).
Simplifying and using the explicit expressions of the angular For the case M2 < 2EI2 (j > 1), we must use the reciprocal
velocity in terms of Jacobi elliptic functions (Eq. (37.10), modulus transformation on the Incomplete Elliptic Integral
Ref. 14 and above Eq. (A1)), we arrive at the expression of of the Third Kind (19.7.4 of Ref. 35, p. 492, or 162.02 p. 39
the time rate of change of / of Ref. 34) as well as the reciprocal modulus transformation
d/ M A for the Jacobi Elliptic Sine function (162.01 p. 39 of Ref. 34)
¼ þ ; (A5) and definition 22.16.1 of Ref. 35 to give
dt I3 B þ C sn2 ð Dtjj2 Þ
with M M ðI 3  I 1 Þ 1
/ðtÞ ¼ tþ P
I3 I1 I3 D j
A ¼ MðI1 þ I3 ÞðI3  I2 Þ < 0; (A6a)  
I 3 ðI 2  I 1 Þ  
B ¼ I1 I3 ðI3  I2 Þ > 0; (A6b)   2 ; am jDtjj2 jj2 :
j I 1 ðI 3  I 2 Þ
(A11)
C ¼ I32 ðI2  I1 Þ > 0; (A6c)
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðI3  I2 ÞðM2  2EI1 Þ As before j is defined in Eq. (A3), and D is given in Eq.
D¼ ; (A6d) (A6d).
I1 I2 I3
Note that in the region M2 < 2EI2, j ¼ 1/k with k defined
and j defined in Eq. (A3). This is valid for all initial through Eq. (16b). Thus, we may rewrite Eqs. (A10) and
conditions. (A11) with notation in the rest of this paper as

8  
> M M ðI 3  I 1 Þ I 3 ðI 2  I 1 Þ   2
>
> tþ P  2
; am Dtjk jk M2 > 2EI2
>
< I3 I1 I3 D I 1 ðI 3  I 2 Þ
/ðtÞ ¼ ! ! (A12)
>
> M M ðI 3  I 1 Þ k 2 I 3 ðI 2  I 1 Þ Dt 2
> 2
M2 < 2EI2 :
: I3 t þ I1 I3 D kP  I1 ðI3  I2 Þ ; am k k jk
>

253 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 253
a) 22
Electronic mail: nmech@vsl.cua.edu L. Bates, R. Cushman, and E. Savev, “The rotation number and the herpol-
1
“Plasma Ben, Dancing T-handle in zero-g, HD,” <https://www.youtube. hode angle in Euler’s top,” Z. Angew. Math. Phys. 56, 183–191 (2005).
23
com/watch?v¼1n-HMSCDYtM>. H. Goldstein, C. Poole, and J. Safko, Classical Mechanics (Addison
2
H. Moffatt, “Euler’s disk and its finite-time singularity,” Nature 404, Wesley, San Francisco, 2002).
24
833–834 (2000). E. T. Whittaker, A Treatise on the Analytical Dynamics of Particles and
3
M. A. Jalali, M. S. Sarebangholi, and M.-R. Alam, “Terminal retrograde Rigid Bodies (Cambridge U.P., Cambridge, UK, 1988).
25
turn of rolling rings,” Phys. Rev. E 92, 032913 (2015). J. W. Zwanziger, M. Koenig, and A. Pines, “Berry’s phase,” Ann. Rev.
4
M. Berry and P. Shukla, “Slow manifold and Hannay angle in the spinning Phys. Chem. 41, 601–646 (1990).
26
top,” Eur. J. Phys. 32, 115–127 (2010). M. S. Ashbaugh, C. C. Chicone, and R. H. Cushman, “The twisting tennis
5
M. Berry and J. Hannay, “Classical non-adiabatic angles,” J. Phys. A 21, racket,” J. Dyn. Differ. Equations 3, 67–85 (1991).
27
L325–L331 (1988). M. Levi, Classical Mechanics with Calculus of Variations and Optimal
6
R. Montgomery, “How much does the rigid body rotate? A Berry’s phase Control: An Intuitive Introduction (American Mathematical Society,
from the 18th century,” Am. J. Phys. 59, 394–398 (1991). Providence, RI, 2014), Vol. 69.
7 28
J. H. Hannay, “Angle variable holonomy in adiabatic excursion of an inte- V. Postell and T. Uzer, “Quantization of the asymmetric top using quan-
grable Hamiltonian,” J. Phys. A 18, 221–230 (1985). tum action-angle variables,” Phys. Rev. A 41, 4035–4037 (1990).
8 29
D. Xiao, M.-C. Chang, and Q. Niu, “Berry phase effects on electronic J. W. Blaker, M. Sidran, and A. Kaercher, Technical Report 472731, RE-
properties,” Rev. Mod. Phys. 82, 1959–2007 (2010). 155, Research Department, Grumman Aircraft Engineering Corp.,
9
R. Resta, “Manifestations of Berry’s phase in molecules and condensed Bethpage, NY, 1962.
30
matter,” J. Phys.: Condens. Matter 12, R107–143 (2000). G. Herzberg, Infrared and Raman Spectra of Polyatomic Molecules (D.
10
C. A. Mead, “The geometric phase in molecular systems,” Rev. Mod. Van Nostrand Company, New York, 1945).
31
Phys. 64, 51–85 (1992). C. H. Townes and A. L. Schawlow, Microwave Spectroscopy (Dover
11
M. V. Berry, “Quantal phase factors accompanying adiabatic changes,” Publications, Mineola, NY, 2013).
32
Proc. R. Soc. London, A 392, 45–57 (1984). S. Frauendorf and F. D€ onau, “Transverse wobbling: A collective mode in
12
J. Anandan, “The geometric phase,” Nature 360, 307–313 (1992). odd-A triaxial nuclei,” Phys. Rev. C 89, 014322 (2014).
13 33
M. Berry, “The geometric phase,” Sci. Am. 259, 46–55 (1988). S. Ødegård, G. B. Hagemann, D. R. Jensen, M. Bergstroem, B. Herskind,
14
L. Landau and E. Lifshitz, Mechanics, Vol. 1 of Course of Theoretical G. Sletten, S. Toermaenen, J. Wilson, P. Tjøm, I. Hamamoto et al.,
Physics (Butterworth Heinemann, Oxford, UK, 1976). “Evidence for the wobbling mode in nuclei,” Phys. Rev. Lett. 86,
15
J. Lawson and M. Rave, “Spacewalks and amusement rides: Illustrations 5866–5869 (2001).
34
of geometric phase,” Math. Mag. 89, 105–121 (2016). P. Byrd and M. Friedman, Handbook of Elliptic Integrals for Engineers
16
S. Gil, “A mechanical device to study geometric phases and curvatures,” and Physicists (Springer-Verlag, Berlin, 1971).
35
Am. J. Phys. 78, 384–390 (2010). F. W. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark, NIST
17
J. Jose and E. Saletan, Classical Dynamics: A Contemporary Approach Handbook of Mathematical Functions Hardback and CD-ROM
(Cambridge University Press, Cambridge, UK, 2000). (Cambridge U.P., New York, 2010).
18 36
J. E. Marsden et al., Lectures on Mechanics (Cambridge U.P., Cambridge, It might be objected that by the time the initial condition (hx and /x )
UK, 1992), Vol. 174. makes it to the x – z plane, the angular velocity may no longer be of unit
19
J. Robbins, “The Hannay angle, thirty years on,” J. Phys. A 49, 431002 magnitude; in fact, in general, it will not). However, since Eq. (20) is inde-
(2016). pendent of jx0 j, we may renormalize and retain the same Da, but not nec-
20
J. B. Hart, R. E. Miller, and R. L. Mills, “A simple geometric model for essarily the same exact dynamics. Equivalently (as discussed in Sec.
visualizing the motion of a Foucault pendulum,” Am. J. Phys. 55, 67–70 III A), we may rescale our unit of time to give jx0 j ¼ 1. This requires scal-
(1987). ing the time by the ratio of the periods. The Binet ellipsoid and angular
21
M. Levi, “Geometric phases in the motion of rigid bodies,” Arch. Ration. momentum sphere are simply magnified or reduced, but the polhode curve
Mech. Anal. 122, 213–229 (1993). shape is maintained.

254 Am. J. Phys., Vol. 87, No. 4, April 2019 Nicholas A. Mecholsky 254

You might also like