Lecture10 PDF
Lecture10 PDF
Alice Quillen
Contents
1 Infinitesimal Transformations and Lie Groups 2
1.1 Lie Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Rotation Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Example: Locomotion of Deformable Jello Cubes . . . . . . . . . . . . . . . 7
1.4 The Jello cube! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5 Mechanical Connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.6 A fibre bundle and a Connection . . . . . . . . . . . . . . . . . . . . . . . . 16
2 Rigid Bodies 17
2.1 Configuration space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Generators of Rotation Matrices . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 The kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4 Example for the Moment of Inertia tensor . . . . . . . . . . . . . . . . . . . 22
2.5 Stability of a Freely Rotating Body . . . . . . . . . . . . . . . . . . . . . . . 22
2.6 Rigid Body motion from a Lagrangian . . . . . . . . . . . . . . . . . . . . . 24
2.7 A Hamiltonian, a moment map . . . . . . . . . . . . . . . . . . . . . . . . . 26
3 Rotational Dynamics 29
3.1 Euler angles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.1 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2 Lagrangian and Hamiltonian in terms of Euler angles . . . . . . . . . . . . . 31
3.2.1 Axisymmetric bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Torque free rotation of the axi-symmetric body . . . . . . . . . . . . . . . . 34
3.4 Euler angles of the torque free axi-symmetric body . . . . . . . . . . . . . . 35
3.5 Frequencies seen by an external viewer - asteroid rotation . . . . . . . . . . 37
3.6 Long axis and short axis modes . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.7 Comments on Damping of Wobbling . . . . . . . . . . . . . . . . . . . . . . 40
3.8 Andoyer-Deprit variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.8.1 Principal axis rotation . . . . . . . . . . . . . . . . . . . . . . . . . . 45
1
3.9 Potential energy of a rigid body from an external force field . . . . . . . . . 45
3.10 MacCullagh’s Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.10.1 For an axisymmetric body . . . . . . . . . . . . . . . . . . . . . . . . 49
3.11 Precession of an axisymmetric body in a circular orbit . . . . . . . . . . . . 49
3.12 Quaternions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.13 A Lie-Poisson numerical method for Solid Body rotation with Quaturnions 57
4 Tidal Evolution 58
4.1 Gravitational Potential Expansion in Legendre Polynomials . . . . . . . . . 58
4.2 Expansion of the gravitational potential of a point mass in spherical harmonics 59
4.3 Tidal deformation of a spherical body . . . . . . . . . . . . . . . . . . . . . 60
4.4 Love numbers for fluid bodies . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.4.1 Love numbers for an incompressible homogeneous elastic body . . . 63
4.5 Rotational Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.6 Tidal spin down . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.6.1 Conventions for delay . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.6.2 Associated orbital drift rate during spin down . . . . . . . . . . . . . 68
4.7 Other timescales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5 Spin Resonances 68
5.1 Types of Spin Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Spin-Orbit resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.3 The dissipative spin-orbit resonance problem . . . . . . . . . . . . . . . . . 71
6 Problems 72
x0 = f (x, a)
2
where a is the r parameters. We desire that for a a vector of zeros, we have the identity
x = f (x, 0, 0, 0....)
where evaluate at aj = 0 for j 6= i. The above f is a matrix or an operator and L has the
dimension of f . Taking into account that f gives a vector we can also write this as
∂fj ∂
Li = (1)
∂ai ai =0 ∂xj
(using summation notation for j) where ∂x∂ j gives a basis for the tangent space, and these
can be used to operate on any function.
Above we have written Li as a differential operator. It operates on a function of x.
With g(x) a function, then Li g(x) is a number so Li g(x) is a function of x. Using two
different differential operators Li , Lj we can operate sequentially on a function Li (Lj g(x)).
Let us consider two vector operators and write them more simply as
∂
A = ai
∂xi
∂
B = bj
∂xj
where ai , bi are functions of x and we have used summation notation. Our generators Li , Lj
are in this form.
∂ ∂
ABg = ai bj g
∂xi ∂xj
∂2g ∂bj ∂g
= ai bj + ai
∂xi xj ∂xi ∂xj
∂2 ∂bj ∂
AB = ai bj + ai
∂xi ∂xj ∂xi ∂xj
3
We notice that the second derivatives cancel. That means that using the commutator our
infinitesimal generators can operate on each other and give us another generator.
The infinitesimal operators form a Lie algebra with commutators (called Lie brackets)
that satisfy
[Li , Lj ] = ckij Lk
with ckij called structure constants. The commutator is a binary operator that satisfies the
Jacobi identity, is antisymmetric and is bilinear w.r.t to multiplication with elements of a
field (typically real or complex numbers).
On a matrix Lie group, we can define an exponential map for any operator in the Lie
algebra, giving us a map between elements of the Lie algebra and the Lie Group. For u in
the Lie algebra
2 3
eu = 1 + u + u2 + u3 ....
2 3!
and this gives an element in the Lie group. In this sense, elements of the Lie algebra are
infinitesimal generators for the group.
u2 v2
eu ev = (1 + u + ....)(1 + v + ....)
2 2
1 1
e(u+v) = 1 + u + v + (u + v)2 .... = 1 + u + v + (u2 + v 2 + uv + vu)....
2 2
We notice that
eu ev 6= eu+v
Locally
u v 1 1 1
e e = exp u + v + [u, v] + [[u, v], v] − [[u, v], u]...
2 12 12
known as the Baker-Campbell-Hausdorff formula.
Consider two near identity operators A, B and alternating the order that the transfor-
mations are done (see Figure 1 and 7). To second order
A2 B2 A2 B2
−A −B A B
e e e e ≈ I −A+ I −B+ I +A+ I +B+
2 2 2 2
A 2 A2 B 2 B 2
= I −A−B+A+B+ + + + + AB − AB − BA + AB − A2 − B 2
2 2 2 2
= I + [A, B]
∼ e[A,B]
The commutator helps us compute what happens when infinitesimal transformation don’t
commute.
4
Figure 1: Alternating infinitesimal transformations need not commute.
is the generator for R(φ). Expanding R(φ) for small φ (to first order in φ)
1 −φ 0
R(φ) ∼ φ 1 0 = I + φL̂z (4)
0 0 1
with
0 −1 0
L̂z = 1 0 0
0 0 0
5
a matrix describing the infinitesimal rotation about the z axis. We can similarly find
0 0 1
L̂y = 0 0 0
−1 0 0
0 0 0
L̂x = 0 0 −1
0 1 0
for rotations about x and y axes. For a small rotation φ about z
x0 = (I + φL̂z )x
0
x 0 −1 0 x x − φy
y 0 = I + φ 1 0 0 y = y + φx
z 0 0 0 0 z z
Suppose we have a function
F (x, y, z)
and we transform it for a small φ.
∂F ∂F
F (x0 , y 0 , z 0 ) = F (x − yφ, y + xφ, z) = F (x, y, z) − y φ+x φ
∂x ∂y
The change in F
∂F ∂F
∆F = −y +x φ
∂x ∂y
We associate the operator
Lz = −y∂x + x∂y (5)
with our infinitesimal transformation. And this is a vector corresponding to the z compo-
nent of angular momentum. So if the coordinate system is rotated by a small amount φ
about the z axis we can compute the change to any function of the coordinates with
∆F = Lz F φ
6
−x sin φ − y cos φ
∂f
= x cos φ − y sin φ
∂φ
0
−y
∂f
= x
∂φ φ=0
0
∂fx ∂fy
Lz = ∂x + ∂y
∂φ φ=0 ∂φ φ=0
= −y∂x + x∂y
consistent with what we have above in equation 5.
This example illustrates two way of describing the infinitesimal operators. One way
with matrices. The other way with differential operators. In both cases we can think of
the infinitesimal operators as lying in the tangent space of the manifold. These operators
also describe the local structure of the Lie group of symmetry transformations.
7
Figure 2: Two masses separated by an actuator that can change its length. If the two
masses have bases with different frictional coefficients and the actuator’s length varies
periodically but not sinusoidally, then the scallop can inch forward. A model explored by
Wagner and Lauga 2013.
Figure 3: The green spring represents an actuator and the green arc an angular spring. By
varying the length of the actuator the horse can be made to jump forwards. Based on the
model by Zhou et al. 2014.
Figure 4: A bristle crawler by Noselli and Desimone. The two legs are flexible and their
angle changes with respect to the rough surface.
8
Figure 5: Variables describing body state for the jello parallelogram.
Dα :
L0 = L(1 + α)
x0 = x
f0 = f (6)
9
The linear transformation of the body itself is chosen to maintain the parallelogram angles.
We assume the transformation does not shift the center of mass. With friction on the base,
this is equivalent to assuming that the transformation is done quickly enough that friction
at the base does not significantly move the center of mass.
Slow shear transformations we describe with
Sβ :
L0 = L
βA
x0 = x +
2L
f0 = f +β (7)
Here A = L20 is the area of the parallelogram face. The length coordinate is preserved but
the angle θ changes. We assume that the transformation is done slowly enough that the
base does not shift during the deformation and as a consequence the transformation causes
a shift in the the horizontal location of the center of mass.
Translations are described with
Tc :
L0 = L
x0 = x + c
f0 = f (8)
q = (x, L, f )
a = (c, α, β)
q0 = G(q, a)
Our transformation functions (the three expressions 6,7, 8) are elements in the group G.
Infinitesimal generators (as operators) or elements in the Lie algebra of G at point q can
be computed using
X ∂Gj ∂
Li =
∂ai ∂qj
j
ai =0
10
Figure 6: Volume preserving deformations of a jello cube with a frictional base. We assume
that the center of mass stays fixed when the jello is stretched vertically. We assume that
the base is held fixed by friction when the jello is is sheared.
11
Figure 7: The deformations don’t commute and their consecutive application translates
the center of mass. A slow shear followed by a stretch, followed by the opposite slow shear
and the opposite stretch causes a translation. The jello cube marches forward.
12
We compute the derivatives
L
∂Dα (q)
= 0 (9)
∂α α=0
0
0
∂Sβ (q) A
= 2L (10)
∂β β=0
1
0
∂Tc (q) 1
= (11)
∂c c=0
0
For the stretches, shears and translations, respectively, the infinitesimal operators at posi-
tion q = (x, L, f ) are
LD = L∂L (12)
A
LS = ∂x + ∂f (13)
2L
LT = ∂x (14)
The space of translations along x is a sub-algebra. The infinitesimal operators are also
vectors in the tangent space of the manifold Q at point q. From the infinitesimal operators
we compute the commutator
A A
[LD , LS ] = LD LS − LS LD = − ∂x = − LT (15)
2L 2L
The operators don’t commute. The commutator gives translation of the center of mass
caused by a series of body transformations. If the operations are done in the order stretch,
shear, inverse stretch, inverse shear, our jello parallelogram should march forward in the
horizontal direction.
The other commutators
[LD , LT ] = [LS , LT ] = 0
The Lie algebra is known as the three-dimensional Heisenberg algebra and it has Bianchi
classification type II.
13
1.5 Mechanical Connection
We can think of configuration space as a principal bundle with full space Q = H × M
and H, the fibre that is also the group of translations of the center of mass, and M the
manifold of body shapes (R2 with coordinates (L, f )). A principal bundle also requires a
projection map π : Q → M that is probably (x, L, f ) → (L, f ).
Body motions are constrained by friction to lie in a horizontal subspace of T Q that is
spanned by the two vectors LD , LS .
A
LD = L∂L LS = ∂x + ∂f (16)
2L
from equation 14. Locally the tangent space of Q can be decomposed into horizontal and
vertical subspaces.
A mechanical connection, Γ is a 1 form (operating on a vector in the tangent space of
Q) that is also a function of elements in the Lie algebra of H. The mechanical connection
sends to zero (projects away) anything in the horizontal subspace. The horizontal sub-
space is that of the body deformation. The vertical subspace is that associated with body
transformation. A mechanical connection is a one form that gives us elements in the Lie
algebra of H, does not change vertical vectors in Q and is zero when applied to vectors in
the horizontal subspace (here spanned by equation 16). We require that the connection Γ
applied to LT give us LT = ∂x and Γ applied to LD and LS give us zero. The mechanical
connection that achieves this is
A
Γ = dx − df ∂x (17)
2L
A
ΓLD = dx − df ∂x on L∂L
2L
A
= −(df L∂L )∂x
2L
= 0
14
A A
ΓLS = dx − df ∂x on ∂x + ∂f
2L 2L
A A
= − ∂x
2L 2L
= 0
Changes in the jello cube body space happen along curves that have tangents specified
by LS and LD . This means they lie in a plane spanned by vectors that are sent to zero by
Γ. A curve that remains in the horizontal space can return to the same f, L coordinates.
But integrating around this loop gives
I
∆x = Γds
where dΓ is a curvature.
The curvature of the connection is
where hu is the horizontal lift of vector u and hv are the horizontal lifts of vector u, v ∈ Tq Q.
Here h essential projects a vector in Tq Q to the tangent space of M (horizontal projection!).
The curvature can be computed from the exterior derivative of the connection. Inte-
grating the connection about a loop is equivalent to integrating the curvature over the area
inside the loop.
With a trajectory given in the horizontal space, the connection can be integrated to
find out how big a change takes place in the vertical space. A loop in the horizontal space
may not close in the vertical space. Integrating the curvature over the area within the loop
in the horizontal space also gives the change that takes place in the vertical space.
More specifically, we write our connection again, (equation 17)
A
Γ = dx − df ∂x (18)
2L
We take the exterior derivative of Γ
A A
dΓ = −( + 1)dx ∧ df + dL ∧ dx + dL ∧ df (19)
2L 2L2
Restricting to the horizontal subspace for L, f we get a curvature
A
Ω= dL ∧ df
2L2
15
We consider a loop in L, f space and use Stokes theorem
I Z Z
Γds = dΓ = Ω
A A
Integrating the connection over the loop is equivalent to integrating the curvature over the
area inside the loop.
16
2 Rigid Bodies
We take our rigid body to be a distribution of particles each with coordinate r with
respect to the center of mass of the body. We let the body spin about an axis Ω. The
velocity of a single particle with mass mi , position ri and velocity vi
vi = Ω × r i
The angular momentum of this particle (measured from the origin) is
Li = mi ri × vi = mi ri × (Ω × ri )
We recall a vector identity
A × (B × C) = (A · C)B − (A · B)C
we can write the angular momentum of the particle
Li = mi ri2 Ω − (ri · Ω)ri
Integrating this up over the entire body the entire angular momentum
Z
L = ρ(r) r2 Ω − (r · Ω)r dV
Using summation notation (and now the indices represent coordinate directions, not a
single particle)
Z
Lj = ρ(r) (xi xi Ωj − xj xk Ωk ) dV
Z
= ρ(r (xi xi δjk − xj xk ) dV Ωk
= Ijk Ωk
17
with moment of Inertia tensor
Z
Ijk ≡ ρ(r) (xi xi δjk − xj xk ) dV
Question Suppose we have a particle distribution and pick a spin value and axis Ω. Each
particle has coordinate xi with respect to the center of mass of the body. We set the initial
velocity vi of each particle i at position xi in the distribution equal to
vi = Ω × xi
(see Figure 8). This question is relevant for generating initial conditions for spinning body
simulations.
Q: Is the initial angular momentum parallel to Ω?
A. The answer is usually no. Angular momentum for a particle is xi ×vi = xi ×(Ω×xi )
and so in general contains a component that is perpendicular to Ω. After summing over
all particles the parts perpendicular to Ω do not cancel unless the particle distribution has
a lot of symmetry.
Q: How do you set the velocity vector of each particle to ensure that the total angular
momentum is a chosen vector L?
A. Compute the moment of inertia tensor from the particle distribution and then com-
pute its inverse. Compute Ω = I−1 L. Use this Ω to set the particle velocities. Because
I is not necessarily diagonal and L not necessarily along a principal axis, L need not be
parallel to Ω.
Rigid body rotation is confusing because two coordinate frames are involved, the inertial
frame and the body frame. The motion of the rigid body is observed in the inertial frame
whereas it is simpler to calculate the equations of motion in the body-fixed principal axis
frame. The body rotes about the angular velocity vector Ω but this is not in the same
direction as the angular momentum vector L. Euler’s equations are given in the body-fixed
frame in which the moment of inertial tensor is diagonal and is constant. The solution then
must be rotated back into the inertial frame to describe the rotational motion as seen by
an external observer.
18
R the trajectory of a rigid body (w.r.t to its center of mass) is R(t). Neglecting external
forces the total energy is equal to the kinetic energy. A Lagrangian description depends on
the kinetic energy with coordinate describing the body orientation at a particular time and
time derivative depending on the time derivative of body orientation which is equivalent
to its spin.
x = Rx0
where R is a rotation matrix relating two reference frames. We can take x0 to be in the
body’s frame and x to be in an inertial frame.
Using the infinitesimal generators from the Lie algebra of the rotation group (SO(3))
19
and these are consistent with equation 21. More generally [lk x] has α component
We use these generators to compute the velocity of a particle in the rotating frame
d
ẋ = (Rx0 ) = Ṙx0
dt
d αi li d
= e x0 = (αi li )eαj lj x0
dt dt
= (α̇ · l)Rx0
v = Ωi li x = (Ω · l)x
where we associate α̇ with Ω. Writing out the dot product (using equation 23)
0 −Ωz Ωy
Ωi li x = (Ω · l)x = Ωz 0 −Ωx x = Ω × x
−Ωy Ωx 0
∂T (Rx0 ) = ∂T (eαi li x0 )
= Ωi li eαj lj x0
= Ωi li Rx0
= Ω i li x
∂
with x = Rx0 and ∂T = ∂t . Note that li x has components
so
or
as expected.
We took a time derivative of coordinate x but we could have done this for any static
vector. Suppose L0 is a static vector.
L = RL0
20
If the vector L0 in one frame is constant then
L̇ = ṘL0 = Ω × L
dL
where L̇ = dt . If our vector L0 is not constant then we would get
L̇ = ṘL0 + RL̇0 = Ω × L + ∂t L (26)
∂ d
where ∂t = ∂t which is not the same as dt .
This looks like an advective derivative. It also
looks a lot like Euler’s equations! In fact if we chose L0 to be in a frame fixed with the
body (for example aligned with the body’s principal axes ) then ∂t L0 = IΩ̇. This follows
as I is constant in this frame. Here L̇ is the torque and we find
I Ω̇ + Ω × (IΩ) = N (27)
with N the applied torque. This gives Euler’s equations.
We take L0 in a frame aligned with a body’s principal axes where the the moment of
inertia matrix is I = diag(Ix , Iy , Iz ),
L0,x = Ix Ωx
L0,y = Iy Ωy
L0,z = Iz Ωz
Equation 27 is
Ix Ω̇x + (Iz − Iy )Ωy Ωz = Nx
Iy Ω̇y + (Ix − Iz )Ωz Ωx = Ny
Iz Ω̇z + (Ix − Iy )Ωy Ωx = Nz
21
Sticking this back into the integral we recognize that
1
T = ΩIΩ
2
where spin Ω is a vector but I is the moment of inertia tensor.
22
where I1 , I2 , I3 are the moments of inertia about the principal axes (eigenvalues of moment
of inertia matrix) and ω1 , ω2 , ω3 are the components of the spin vector in the principle axes
directions (eigenvector directions of the moment of inertia matrix). We can label the axes
so that I1 > I2 > I3 and start the body rotating nearly about direction x̂2 with
ω = α1 x̂1 + sx̂2 + α3 x̂3
with s large and α1 , α2 1 very small perturbations. We put this into Euler’s equations
(equations 28)
(I2 − I3 )
ω̇1 = sα3
I1
(I3 − I1 )
ω̇2 = α1 α3 ∼ 0
I2
(I1 − I2 )
ω̇3 = sα1
I3
The middle equation is second order in small quantities so ω̇2 ∼ 0. If we allow α1 , α2 to be
functions of time
(I2 − I3 )
α̇1 = sα3
I1
(I1 − I2 )
α̇3 = sα1
I3
and putting this together
(I2 − I3 ) (I1 − I2 ) 2
α̈1 = s α1
I1 I3
(I2 − I3 ) (I1 − I2 ) 2
α̈3 = s α3
I1 I3
If the factor on the right is positive then the solutions are exponential growing or shrinking.
If the factor on the right is negative then the solutions are oscillating. If the quantity
(I2 − I3 )(I1 − I2 )
is positive then we can have exponentially growing values for ω1 , ω2 . Since I1 > I2 > I3
this quantity is positive and the system is unstable and will quickly start to spin about a
different direction than started.
Redoing this analysis starting with spin about x̂1 the relevant quantity that must
be negative for stability is (I1 − I3 )(I2 − I1 ) and this is negative with our order for the
moments. Redoing the analysis starting with spin about x̂3 the relevant quantity that must
be negative for stability is (I3 − I1 )(I2 − I3 ), again negative with our ordering.
The body can stably spin about the axis of smallest moment or about the axis of the
largest moment but not the intermediate one.
23
2.6 Rigid Body motion from a Lagrangian
We notice that the kinetic energy is in fact a Lagrangian density
Z
dV 1
L=T = ρ (∂T (Rx))2 = ΩIΩ (29)
2 2
where the kinetic energy is integrated over the body volume. Again x is in the body frame
and Rx is in the inertial frame. The Lagrangian itself is integrated over a path
Z t2
L= L(t)dt
t1
with path R(t) specified by how the rotation varies in time. The rotation matrices are
functions of angles, α that specify body orientation. Each rotation is a point inside SO(3).
The body rotations so α(t), these angles are functions of time. The time derivatives of
these angles are α̇ = Ω a spin vector. The spin vector is a member of the tangent space of
a point in SO(3). Our Lagrangian density is a function
Taking the angles and their time derivatives we can compute Lagrange’s equations.
∂L
= IΩ (30)
∂Ω
which is the angular momentum. Using equation 29
d ∂L
= L̇ = IΩ̇ (31)
dt ∂Ω
Now for the hard part, taking the derivative of the Lagrangian with respect to the
angles for the rotation (the coordinate in SO(3)). Using equation 20 to write the rotation
matrix in terms of α
∂ ∂ αj lj
Rx = e x
∂αk ∂αk
= lk eαj lj x
= lk Rx
= lk y (32)
24
so if we dot this with another vector z
∂y
z· = −kβγ yγ zβ
∂α
= y×z
v = ∂T (Rx)
= Ω × Rx
= Ω×y
∂ v2
= Ω2 y · (lk y) − (Ω · y)Ωi lk yi
∂αk 2
∂ v2
= Ω2 y × y − (Ω · y)(y × Ω)
∂α 2
= (Ω · y)(Ω × y)
= (Ω · y)v (34)
Ω × L(y) = Ω × (y × v)
= (Ω · v)y − (Ω · y)v
= −(Ω · y)v (35)
Ω · v = Ω · (Ω × y) = 0.
25
∂L
= −Ω × L (36)
∂α
This and equation 31 with Lagrange’s equations give
L̇ + Ω × L = 0.
Using the moment of inertia tensor in a frame fixed with the body we get Euler’s equations.
I Ω̇ + Ω × IΩ = 0.
Av = Ω × v
26
Figure 9: Euler angles. The green line is the intersection of the i, j and I, J planes and is
called a line of nodes with direction n̂. It is computed from the vector K × k. The angle
θ is a nutation angle. The angle φ is a spin precession angle (as seen from the external or
fixed reference frame). ψ is sometimes called a precession angle as seen from the body. ψ̇
rotates the ij axes in the ij plane with respect to a fixed line of nodes. φ̇ rotates the line
of nodes n̂.
27
Figure 10: A toy plane! Practical uses for Euler angles include study of plane stability.
28
3 Rotational Dynamics
3.1 Euler angles
Following Celleti’s book:
We start with a fixed point O and a reference frame i, j, k that coincides with a body’s
principle axes (O, i, j, k). Now we chose an inertial frame with the same origin and three
different axes (O, I, J, K). Let n be the line of nodes defined by the intersection of the i, j
and I, J planes. The direction
K×k
n̂ =
|K × k|
The Euler angles θ, φ, ψ are as follows:
• 0 ≤ θ ≤ π is the nutation angle and is that between K and k axes.
• 0 ≤ ψ ≤ 2π is the proper rotation angle and is that between n̂ and i axis.
• 0 ≤ φ ≤ 2π is the precession angle and is that between n̂ and I axis.
The angle ψ is also sometimes called a precession angle as it can describe the precession
of the spin and angular momentum vectors in the body’s frame.
We need to relate variations in the Euler angles to the spin in the body frame.
ωi = φ̇ sin θ sin ψ + θ̇ cos ψ
ωj = φ̇ sin θ cos ψ − θ̇ sin ψ
ωk = φ̇ cos θ + ψ̇ (37)
The three rotations described below allow us to derive these expressions.
Looking at figure 9 we examine the angles. With θ, φ constant ψ̇ gives a rotation of
the i, j plane with respect to the line of nodes. The body is essentially spinning around
the k body axis. The angle φ moves the line of nodes with respect to the inertial frame. If
the body spin is oriented along k we can look at its projection into the I, J plane. If the
projected spin axis advances we say the body spin precesses. Taking k to be the spin axis,
the angle between it and the K axis we refer to as a nutation angle, θ.
• The angle θ is that between k (body principal axis) and K (inertial frame axis).
cos θ = k · K.
• The angle ψ is that between the body principal axis i and the line of nodes n̂
cos ψ = i · n̂.
29
3.1.1 Rotations
Starting in the inertial frame (xyz), three rotations transform into the body frame (x0 y 0 z 0 ).
1. A counter-clockwise rotation by φ about K or the current z axis.
2. A counter-clockwise rotation by θ about the current x axis (here the x000 axis).
3. A counter-clockwise rotation by ψ about the current z axis (here the z 0 or z 000 axis).
First rotate by the precession angle φ then tilt by the nutation angle θ then rotate
by ψ.
We start with xyz in inertial frame, rotate by φ giving x00 , y 00 , z 00 , rotate by θ giving
x000 , y 000 , z 000 then rotate by ψ giving x0 , y 0 , z 0 in body frame.
00
x cos φ sin φ 0 x
y 00 = − sin φ cos φ 0 y
z 00 0 0 1 z
x000
00
1 0 0 x
y 000 = 0 cos θ sin θ y 00
z 000 0 − sin θ cos θ z 00
x0
000
cos ψ sin ψ 0 x
y 0 = − sin ψ cos ψ 0 y 000
z0 0 0 1 z 000
Angular rotation rates arise from time derivatives of each rotation;
ωφ = φ̇ẑ
ωθ = θ̇x̂00
ωψ = ψ̇ẑ0 (38)
Combining these (equations 38 and 39) together we get equations 37 for the body frame
(x0 , y0 , z0 ) repeated here
30
All together the three rotations give this transformation
x0 = (cos ψ cos φ − sin ψ sin φ cos θ)x + (cos ψ sin φ + sin ψ cos φ cos θ)y + sin ψ sin θz
y 0 = −(sin ψ cos φ + cos ψ sin φ cos θ)x + (− sin ψ sin φ + cos ψ cos φ cos θ)y + cos ψ sin θz
z 0 = sin θ sin φx − sin θ cos φy + cos θz, (41)
x = (cos φ cos ψ − sin φ sin ψ cos θ)x0 − (cos φ sin ψ + sin φ cos ψ cos θ)y 0 + sin φ sin θz 0
y = (sin φ cos ψ + cos φ sin ψ cos θ)x0 + (− sin φ sin ψ + cos φ cos ψ cos θ)y 0 − cos φ sin θz 0
z = sin θ sin ψx0 + sin θ cos ψy 0 + cos θz 0 . (42)
• Likewise projecting body z 0 axis into the inertial x, y plane gives φ, with
We use the function atan2 so that the quadrant for the angle is specified.
31
or
With these canonical momenta we can construct a Hamiltonian from the kinetic energy
term and using Euler angles and their associated momenta as canonical variables
H(φ, θ, ψ; pφ , pθ , pψ )
giving us formula for ω1 , ω2 , ω3 in terms of our canonical momenta and angles only. These
are also the angular momentum vector in the body frame. We can replace the spins
ω1 , ω2 , ω3 in the Lagrangian with expressions that depend only on momenta (pθ , pφ , pψ )
and angles giving our Hamiltonian;
2
1 (pφ − pψ cos θ)
H(φ, θ, ψ; pφ , pθ , pψ ) = sin ψ + pθ cos ψ +
2I1 sin θ
2
1 (pφ − pψ cos θ)
cos ψ − pθ sin ψ +
2I2 sin θ
1 2
p
2I3 ψ
With I1 = I2 this simplifies. We note that the choice of angles and momenta is poor
when θ ∼ 0. In that setting it is a good idea to change coordinate system via canonical
transformation.
32
and the Hamiltonian is
1 (pφ − pψ cos θ)2
2 1 2
Hsym (φ, θ, ψ; pφ , pθ , pψ ) = + pθ + p +V (45)
2I1 2
sin θ 2I3 ψ
With V independent of φ, ψ, the Hamiltonian is only dependent on a single angle θ so
momenta pφ , pψ are conserved quantities as is H itself.
In the axisymmetric setting
pφ = (I1 sin2 θ + I3 cos2 θ)φ̇ + I3 ψ̇ cos θ (46)
pψ = I3 (φ̇ cos θ + ψ̇) (47)
pθ = I1 θ̇ (48)
and as long as the potential is independent of φ, ψ, these momenta pφ , pψ remain conserved.
It may be useful to write precession rates in terms of pφ , pψ
(pφ − pψ cos θ)
φ̇ = (49)
I1 sin2 θ
pψ pψ (pφ − pψ cos θ)
ψ̇ = − φ̇ cos θ = − cos θ (50)
I3 I3 I1 sin2 θ
With θ fixed we get constant precession rates, φ̇, ψ̇. The conserved quantity pψ is the
angular momentum projection along the body-fixed principal axis, whereas pφ is the angular
momentum about the z inertial axis. This is probably understood from the last equation
for z in the transformation 42 and applying it to the angular momentum as seen in the
body frame.
What is not obvious that we should be able to find solutions with fixed θ in all axisym-
metric situations. This is equivalent to choosing an orientation for using the Euler angles
– in other words choosing an orientation for the inertial coordinate system.
For torque free rotation we can choose our inertial frame to be with z aligned with the
angular momentum so L = Lẑ that is fixed in the inertial frame. Using our transformation
in equations 41 we find that in the body frame the angular momentum is
Lbody = L(sin ψ sin θx̂0 + cos ψ sin θŷ0 + cos θẑ0 )
and using our relations for ω1 , ω2 , ω3 (equations 37)
L sin ψ sin θ = I1 ω1 = I1 (φ̇ sin θ sin ψ + θ̇ cos ψ)
L cos ψ sin θ = I1 ω2 = I1 (φ̇ sin θ cos ψ − θ̇ sin ψ)
L cos θ = I3 ω3 = I3 (φ̇ cos θ + ψ̇) (51)
Taking the first two equations
L sin θ(sin ψ cos ψ − cos ψ sin ψ) = 0
= I1 ω1 cos ψ − I1 ω2 sin φ = pθ
= I1 θ̇
33
So our choice of coordinate orientation, assuming an axisymmetric body and a constant
angular momentum gave us θ̇ = 0.
for spin ω in the body’s reference frame. We notice that ωz is fixed. Set
Ik − I⊥
Ω= ωz
I⊥
and
ωz = ω cos α.
The angle α is the angle between spin vector ω and body symmetry vector ẑ. The equations
of motion become
A solution is
34
It is convenient to define an angle θ between L and body principal axis ẑ that satisfies
q
L2x + L2y I⊥
tan θ = = tan α (64)
Lz Ik
Consider a moment when ω and L are in the yz plane and so all have zero x components.
In order for ωx = 0 at this time, ψ = 0 and
The first relation follows from equation 65 and the second one from the length of ω with
ωx = 0. Equation 66 gives a relation between θ, α, ω, φ̇ that is satisfied at all times.
35
Combined with equation 64 we find
ω sin α
φ̇ =
sin θ
!1
Ik2 2
2
= ω sin α 2 cot α + 1
I⊥
#1
Ik2
" !
2
2
= ω 1+ 2 − 1 cos α (67)
I⊥
" !#− 1
2
I⊥ 2
= ω 1 + cos2 θ −1 (68)
Ik2
Using ωz = w cos α and the last of equation 65, some further manipulation using equation
64 gives
The last step we have used equation 60. Here we associate ψ̇ with the precession rate of
the angular momentum and spin in the body frame.
For small α (and simultaneously small θ)
Ik
lim φ̇ → ω (70)
α→0 I⊥
Ik
ψ̇ → ω 1 − (71)
I⊥
Ik
ψ̇ I⊥
→ 1− (72)
φ̇ Ik I⊥
Whereas −ψ̇ is the precession rate in the body frame, φ̇ is the spin precession rate in
the inertial frame. The two are not equal. For a prolate body Ik < I⊥ and vice versa for an
oblate body. Ω < 0 for elongated (prolate) bodies but Ω > 0 for flattened (oblate) bodies,
whereas φ̇ is always positive. Precession in the inertial frame is in the same sense as Lz ,
the vertical component of angular momentum. For a prolate body and small α, the rate
φ̇ φ̇
ω < 1, whereas for an oblate body ω >√1.
For a thin disk Ik = 2I⊥ and φ̇ = ω 1 + 3 cos2 α. For small α, the precession rate φ̇ ∼
2ω and Ω ∼ ω and giving the famous 2:1 ratio discussed by Feynman (though apparently
he had which frequency was smaller backwards).
36
We have so far explored the torque-free axisymmetric top starting from Euler’s equa-
tions. We should be able to find the same solution using Hamilton’s equations or Lagrange’s
equations. A difficulty is that we need to chose the inertial frame so that the Euler angle θ
is fixed. Once that is done conserved momenta give the precession rates. I think choosing
the good inertial frame is equivalent to computing the angular momentum direction and
choosing the z axis (for the Euler angles) to align with it. Another inconvenience with the
Euler angles is that we are left without a nice expression for the angular momentum in the
inertial frame.
ez 0 (t) = sin φ sin θex − cos φ sin θey + cos θez . (73)
ez 0 (t) = sin(φ̇t) sin θex − cos(φ̇t) sin θey + cos θez (74)
as φ̇ (equation 67) is constant for a body of revolution. Likewise we can use equation 42
to determine the other two body axes as a function of time as seen in the inertial frame
37
viewer will see two frequencies in the motion of the precessing body, that set by Ω and
that set by φ̇. Two frequencies present in the light curve gives evidence of non-principal
axis rotation.
It is not easy to determine the body axis ratios from the light curve because the bright-
ness depends on illumination angle. In other words, how much sun light is reflected depends
on the angles between the sun to object line, object to viewer line and surface normal vec-
tors. The surface could also have color or albedo variations (spots!).
If the body is triaxial rather than a body of revolution, then θ is not fixed and there
will be an additional frequency present in the motion, called nutation. And there may be
aperiodicity as well in φ̇. For triaxial bodies the spins and some periods for the Euler angles
can be written in terms of elliptic functions. These give two periods and an associated
precession rate that could be evident in a light curve.
Here we defined the Euler angles with respect to the body symmetry axis. In general
you could define them with respect to a long or short principal body axis. This gives two
conventions for specifying them.
To interpret asteroid rotation, you start with spin and angular momentum vectors and
three moments of inertia. Solve Euler’s equations in the body frame. Then compute the
frequencies associated with the Euler angles and compare them to frequencies that are seen
in a periodogram of a light curve.
38
s
2E(Is − U )
Ωl = dnτ (81)
Il (Is − Il )
s
2E(U − Il )
Ωi = snτ (82)
Ii (Ii − Il )
s
2E(U − Il )
Ωs = cnτ (83)
Is (Is − Il )
It turns out that two of the angles are periodic in τ , but the last one is not, though the
rate φ̇s is. The period
s Z π/2
Il Ii Is du
PψS = 4 p (87)
2E(Ii − Il )(Is − U ) 0 1 − k 2 sin2 u
L2
Ii < ≤ Is (88)
2E
s
2E(Is − Ii )(U − Il )
τ =t (89)
Il Ii Is
(Ii − Il ) (Is − U )
k2 = (90)
(Is − Ii ) (U − Ii )
39
3.7 Comments on Damping of Wobbling
If an asteroid is not rotating about a principal axis, then Ω is time dependent. Let a0 , v0 , r0
be the acceleration, velocity and position in the body frame and a, v, r be those in the
inertial frame. To relate the time derivative of a position in the inertial frame to that in
the body frame
d ∂
= +Ω×. (91)
dt ∂t
Apply this to position r0 in the body
dr ∂r0
= + Ω × r0
dt ∂t
v = v 0 + Ω × r0
If the body does not flex much then a0 and v0 (in the body frame) are small and
a ≈ Ω̇ × r0 + Ω × (Ω × r0 ) (92)
The time dependent stress due to the spin variations can be estimated using this equation.
For a prolate or oblate body, Ω depends on a single frequency and this sets the timescale
of elastic stress variations in the body. If the body is triaxial then an entire spectrum of
frequencies is involved. This is relevant as internal friction due to the elastic deformation
of the body may be frequency dependent.
While angular momentum is conserved, the total energy is not due to dissipation or
internal friction inside the elastic body which goes into heat that can be radiated away.
The rate of energy loss is computing by integrating up over the volume of the body the
stress times the strain rate.
40
Figure 11: Andoyer-Deprit angles. The green lines show lines of nodes. The body’s spin
angular momentum is aligned with i3 . The p̂ line of nodes is the intersection of the i, j
plane (body) with the i1 , i2 (spin) plane. The m̂ line of nodes is the intersection of the
I, J plane (inertial) with the i1 , i2 (spin) plane. The n̂ line of nodes is the intersection
of the i, j plane (body) with the I, J plane (inertial). The angles θ, φ, ψ are Euler angles.
The angles g, h, l are Andoyer-Deprit angles. The angles J, K are used in the definition
of Andoyer-Deprit momenta. Under principal axis rotation J = 0, K = θ, m̂ = n̂ = p̂,
φ = h, g = 0, and l = ψ.
41
• (O, I, J, K) is an inertial reference frame (for example defined with respect to a
Laplace plane or the elliptic). Here O refers to the center of mass of our spinning
body.
• (O, i1 , i2 , i3 ) is a spin frame, and i3 is aligned with the angular momentum vector.
Because we have three frames we need more than one line of nodes (see Figure 11) .
• The n̂ line of nodes is the intersection of the i, j plane (body) with the I, J plane
(inertial).
n̂ = K × k/|K × k|
We used this line of nodes when defining the Euler angles.
• The m̂ line of nodes is the intersection of the I, J plane (inertial) with the i1 , i2 plane
(body).
m̂ = i3 × K/|i3 × K|
• The p̂ line of nodes is the intersection of the i, j (body) plane with the i1 , i2 plane
(spin).
p̂ = k × i3 /|k × i3 |
The signs on the cross products need to be checked for convention here! I have adopted
the order in Celleti’s book.
Euler angle lists
• (θ, φ, ψ) are Euler angles of the body frame with respect to the inertial frame.
• (J, g, l) are Euler angles of the body frame with respect to the spin frame.
• (K, h, 0) are Euler angles of the inertial frame with respect to the spin frame, but
assuming that i3 lies on line of nodes m̂.
What are i1 , i2 aligned with? Their directions are not needed to define the Andoyer-
Deprit variables. The angles depend on the lines of nodes rather than the orientation of i1
or i2 .
Andoyer-Deprit variables are a canonical set (G, g), (L, l), (H, h) of momenta paired
with angles.
If M0 is the angular momentum vector M0 = M0 i3 .
G ≡ M0 · i3 = M0 (93)
L ≡ M0 · k = G cos J (94)
H ≡ M0 · K = G cos K (95)
42
The angle K is an obliquity angle between angular momentum vector and inertial
frame axis. It is not one of the Andoyer-Deprit variables. The angle K is that between i3
(angular momentum vector) and K (inertial frame axis).
cos K = i3 · K
The angle J is called the non-principal rotation angle. When it is zero, J = 0, the
body is rotating about a principal body axis. It is not one of the Andoyer-Deprit variables.
The angle J is that between i3 (angular momentum vector) and k (body principal axis).
cos J = i3 · k
The Andoyer-Deprit momenta in terms of those associated with the Euler angles are
pφ = H
pψ = L
pθ = G sin J sin(l − ψ) (96)
Here the Euler angles are defined with respect to the inertial reference frame I, J, K.
The variables G, J, L, l are related to the angular momentum orientation as seen in the
body frame
I1 ω1 = G sin J sin l
I2 ω2 = G sin J cos l
I3 ω3 = L (97)
cos g = m̂ · p̂
43
• The angle h is that between I and line of nodes m̂.
cos h = I · m̂
• The angle l is that between i (body principal axis) and line of nodes p̂.
cos l = i · p̂
Using the relation between Euler angles and associated momenta and the Andoyer-
Deprit variables, the Hamiltonian can be written in terms of the Andoyer-Deprit variables
2
sin l cos2 l 1 L2
1 2 2
H(G, L, H; g, l, h) = (G − L ) + + (98)
2 I1 I2 2 I3
It is independent of H, g, h, so their canonical counter parts h, G, H must be conserved.
A useful relation
G2 − L2 L2
Haxi = +
2I1 2I3
and
G 1 1
ġ = l˙ = − L ḣ = 0 (99)
I1 I3 I1
The body rotates around the symmetry axis with constant velocity and that it precedes
uniformly.
With Euler angles, the solution of precessing axisymmetric body is not obvious (you
must establish that you can find a coordinate system with θ̇ = 0) using Andoyer-Deprit
variables, it is clearer that a rigid body rotates about the symmetry axis with a constant
velocity and precessing with a constant velocity.
The Andoyer-Deprit variables are not well defined if J = 0 or K = 0. However one can
transform (canonically) to these variables
λ1 = l + g + h Λ1 = G
λ2 = −l Λ2 = G − L = G(1 − cos J)
λ3 = −h Λ3 = G − H = G(1 − cos K)
44
The new Hamiltonian is
(Λ1 − Λ2 )2
H(Λ1 , Λ2 , Λ3 ; λ1 , λ2 , λ3 ) =
2I3
sin2 λ2 cos2 λ2
1 2 2
+ Λ1 − (Λ1 − Λ2 ) + (100)
2 I1 I2
It is independent of Λ3 , λ1 , λ3 so their canonical counterparts λ3 , Λ1 , Λ3 must be conserved.
Because the force field is external ∇2 Φ = 0 so Φ,ii (x0 ) = 0. The trace of the potential’s
Hessian is zero because the external potential must satisfy Laplace’s equation.
The total potential energy we find by integrating over body coordinates
Z
3 i 1 i j
U = d yρ(y) Φ(x0 ) + Φ,i (x0 )y + Φ,ij (x0 )y y + ...
2
Z Z
1
= Φ(x0 )m + Φ,i (x0 ) dy ρ(y)yi + Φ,ij (x0 ) dy 3 ρ(y)y i y j
3
2
Z
1
= Φ(x0 )m + Φ,ij (x0 ) dy 3 ρ(y)y i y j
2
45
The first order term drops out because we are using the center of mass coordinate frame.
The zero-th order term is that for a point mass. The interesting term is the second order
term. Because the trace of the Hessian is zero
Z
0 = Φ,ii (x0 ) dy 3 ρ(y)y k y k
Z
= Φ,ij (x0 )δij dy 3 ρ(y)y k y k
We will use this to rewrite the potential energy with the moment of inertial tensor
Z
1
U2 = Φ,ij (x0 ) dy 3 ρ(y)y i y j
2
Z
1 h i
= Φ,ij (x0 ) dy 3 ρ(y) y i y j − δij y k y k
2
1
= − Φ,ij (x0 )Iij
2
Rewriting this the potential term associated with body torque is
1
U2 = − Φ,ij (x0 )Iij (104)
2
If we require more accuracy we we could add in higher order terms, depending upon higher
order moments of both body and potential.
The external force field could be from an oblate planet, a binary or a point mass. In
all these settings we could compute the derivatives of the external potential and from that
the potential energy integrated over the body. The potential energy of the body can be
put immediately into a Lagrangian or a Hamiltonian using one of the choices above for
canonical variables. By taking derivatives of the potential energy function the torque on
the body can be computed.
46
The second order term is the one that gives us our Hessian matrix.
1
3 cos2 ψ − 1
P2 (cos ψ) =
2
1 02 02 02 02
= 3z − x − y − z
2r02
1
2z 02 − x02 − y 02
= 02
2r
where I have oriented the prime coordinate system so that the point mass M lies along z
axis. In this coordinate system
GM
Φ,xx = Φ,yy =
r3
2GM
Φ,zz = − = −2Φ,xx
r3
Taking only second order terms
GM
Φ2,ij = (δij − 3δiz δjz )
r3
Putting this into equation 104 the left hand part will give us the trace of the moment of
Inertia matrix.
GM
U2 = − 3 (trI − 3Izz ) (106)
2r
where Izz is the moment of inertia of m about the axis between our body m and the
external point mass M . This is equivalent to MacCullagh’s formula.
A small non-round body in the gravitational field of a distant point source has a po-
tential energy perturbation given by this (remember to add it to the zeroth order part
−GM m/r). This can be used to compute the torque on m by M which is the same but
opposite to the torque on M by m. A useful formula for tidal computations.
MacCullagh’s formula is often written
47
the non-round body on a point mass, rather than integrating the effect of a point mass on
an extended body.
By taking derivatives of the potential function we can compute the torque. Taking
i, j, k to be unit vectors lying along principal axis of the extended spinning body that have
moments of inertia A < B < C and r is the vector from the center of mass of the extended
spinning body to a point mass M . We can write r in a basis aligned with our principal
body axes
r = r((r̂ · i)i + (r̂ · j)j + (r̂ · k)k
and equation 108 as
I = (r̂ · i)2 A + (r̂ · j)2 B + (r̂ · j)2 C
and equation 107 as
48
3.10.1 For an axisymmetric body
MacCullagh’s formula is usually described as the gravitational potential of an oblate body,
like the Earth. For an oblate body oriented with z along the pole, Ixx = Iyy = I⊥ . This is
often written with I⊥ as the moment of inertia about an axis perpendicular to the body’s
axis (x or y) of symmetry and Izz = Ik is that about an the axis parallel to the axis of
symmetry (here z).
In a coordinate system with I diagonal
The moment I only depends on a latitude angle θ. Let us use a convention θ = 0 gives
the axis of symmetry and θ ∈ [0, π] with θ = π/2 at the equator with z = r cos θ. Taking
n̂ = (cos φ sin θ, sin φ sin θ, cos θ)
Gm G(2I⊥ + Ik − 3I)
U (r, θ) = − −
r 2r3
Gm G(Ik − I⊥ )(1 − 3 cos2 θ)
= − −
r 2r3
Gm G(I k − I ⊥ )
= − + P2 (cos θ) (114)
r r3
The total potential energy of an axisymmetric extended body described with Ik , I⊥ inter-
acting with a point mass M is
GM m GM (Ik − I⊥ )
E(r, θ) = − + P2 (cos θ) (115)
r r3
where r is the distance between M and the center of mass of m and θ gives the angle
between the vector joining the two masses and the axis of symmetry of the axisymmetric
body. In other words cos θ = r̂ · k with k along the body’s axis of symmetry.
For moments of inertia A = B, the torque in equation 112 simplifies to
3GM
T= (C − A)(r · k)(r × k). (116)
r3
49
we can chose a coordinate system oriented with orbit angular momentum aligned in the z
direction,
r = r(cos λt, sin λt, 0)
with λ̇ = n the mean motion. It is convenient to notice that GM/r3 = n2 is the mean
motion as long as the point mass is much larger than that of the precessing body.
For principal axis rotation the body’s spin ŝ = k and with θ the obliquity and φ the
precession angle projected onto the orbital plane,
3n2 (C − A)
hTi = sin θ cos θ (sin φ, − cos φ, 0)
2
is in a direction consistent with φ but not θ varying.
dL
= Cω sin θφ̇ (− sin φ, cos φ, 0).
dt
Equating these two expressions we find a precession rate,
3 (C − A)n2
φ̇ = − cos θ (117)
2 Cω
The equations of motion are consistent with a Hamiltonian system with canonical momen-
tum p = 1 − cos θ and canonical coordinate φ
αs αs
H(p, φ) = cos2 θ = (p − 1)2 (118)
2 2
and
3 (C − A)n2
αs ≡ . (119)
2 Cω
A planet can experience torques (and associated obliquity variations) due to perturba-
tions on its orbit. In this case the torque from the Sun is modulated due to orbit orientation
variations, typically described by variations in the longitude of the ascending node. Bill
Ward and collaborators wrote a series of papers accounting for obliquity variations with this
50
type of model, which is usually called ‘secular spin resonance’. The dynamics is considered
secular as it can be approximated by averaging over both orbit and spin angles.
How does a planet precess if it has moons? Moons can orbit a planet quickly compared
to the precession rate orbit of the planet. If the satellites nodal recession due planetary
oblateness is faster than the planet’s spin precession rate, then the moon’s orbit can be
considered a ring that slowly moves along with the planet and its orbit stays oriented
about the equator of the planet. An adiabatic invariant is preserved. The planet and its
moons are considered to precess together as if they were a single object. Moons affect the
planet’s precession rate as they add to the total spin angular momentum and make the
planet behave as if it were flatter. Saturn would precess more slowly if it did not have a
ring and moon satellite system.
The Moon’s orbit nodal recession period is 19 years and is much faster than the Earth’s
spin precession period which is 26,000 years. However precession of the Moon’s orbit is due
to the torque from the Sun, rather from Earth’s oblateness and the Moon’s orbit precesses
about the normal of the Earth’s orbit rather than about the spin axis of the Earth. The
transition between precession due to Earth’s oblateness and about the Earth’s spin axis
and precession due to the Sun and about the Earth’s orbital normal is called the Laplace
transition. A lock between the orbit nodal precession rate and the spin precession rate is
called a Cassini state. Tidally locked moons tend to be in a Cassini state.
Does the Moon’s inclination with respect to the ecliptic or with respect to the Earth’s
equator remain constant as the Earth precesses? The Moon’s orbit w.r.t. ecliptic has
inclination 5.15 degrees. The Moon’s obliquity (angle between orbit planet and spin axis
is 6.688 degrees). Both angles remain fixed as Earth’s spin axis precesses. The Moon does
affect the precession rate of the Earth, but the oblateness of the Earth no longer affects
the Moon’s orbit because the Moon is too far away.
3.12 Quaternions
We define a quaternion as
q = q0 + qx i + qy j + qz k
where q0 is called the scalar part and (qx , qy , qz ) is called the vector part of the quaternion.
Multiplication of quaternions is performed using the Hamilton product. Using the rules
i2 = j2 = k2 = −1
ij = k = −ji
jk = i = −kj
ki = j = −ik, (120)
51
the Hamilton product of q1 = a1 + b1 i + c1 j + d1 k and q2 = a2 + b2 i + c2 j + d2 k is
q1 q2 = a1 a2 − b1 b2 − c1 c2 − d1 d2
+ (b1 a2 + a1 b2 + c1 d2 − d1 c2 )i
+ (a1 c2 − b1 d2 + c1 a2 + d1 b2 )j
+ (a1 d2 + b1 c2 − c1 b2 + d1 a2 )k. (121)
Notice that θ → −θ and u → −u gives the same rotation. This means that q → −q gives
the same rotation. Flipping the sign of θ or u (but not both) gives q−1 .
For a unit quaternion q = q0 + qx i + qy j + qz k with q02 + qx2 + qy2 + qz2 = 1, the axis of
rotation and angle
(qx , qx , qz ) (qx , qx , qz )
u= q = p (123)
qx2 + qy2 + qz2 1 − q02
q
θ = 2 atan2 qx2 + qy2 + qz2 , q0
q q
= 2 atan2 qx + qy + qz , 1 − (qx2 + qy2 + qz2 )
2 2 2
q
= 2 atan2 2
1 − q0 , q0 . (124)
p0 = qpq−1 . (125)
To carry out consecutive rotations, quaternions can be multiplied using quaternion multi-
plication.
Taking unit quaternion q = q0 + qx i + qy j + qz k, the rotated vector can be written in
terms of a rotation matrix R with
p0 = Rp
and
1 − 2(qy2 + qz2 ) 2(qx qy − qz q0 ) 2(qx qz + qy q0 )
52
To compute RT = R−1 we need only flip the sign of the last term. Rotating a vector p by
q gives
ψ = atan2(Rxz , Ryz )
φ = atan2(Rzx , Rzy )
with order reversed if the quaternion transfers from body frame to inertial frame.
Consider two rotations giving a third RC = RB RA
α α
qA = cos + A sin (128)
2 2
β β
qB = cos + B sin (129)
2 2
γ γ
qC = cos + C sin (130)
2 2
with unit vectors A, B, C and angles α, β, γ. The result is known as Rodrigues formula for
composite rotation and is
γ α β α β
cos = cos cos − sin sin B · A (131)
2 2 2 2 2
γ β α α β β α
sin C = sin cos B + sin cos A + sin sin B × A (132)
2 2 2 2 2 2 2
−1
γ α β β α α β
tan C = 1 − tan tan B · A tan B + tan A + tan tan B × A . (133)
2 2 2 2 2 2 2
For describing rotations, quaternions have some advantages. Quaternions avoid a phe-
nomenon called gimbal lock which can result in pitch/yaw/roll rotational systems. They
ensure rotations are possible in any orientation. As they depend only on 4 numbers, the
notation is more compact than using rotation matrices. From a quaternion representation
of rotations it is easy to read off the axis and angle of rotation (much easier than when
using Euler angles).
The unit quaternion Lie group (S 3 ; 3D sphere in 4 dimensions) is not exactly the
same as the group of rotations SO(3) (special orthogonal matrices). This is because unit
quaternions q and −q give the same rotation.
53
Question Is there any intuition gained by using canonical variables based on quaternions?
54
where r is a vector for each position in the body. Note, that Ω is not equivalent to the
time derivative of a quaternion describing current body orientation. Instead it describes a
quaternion giving the infinitesimal spin rotation associated with rotation.
It may be useful to use the vector identity
(Ω × r)2 = Ω2 r2 − (Ω · r)2
d
(Ω × r)2 = 2Ωi r2 − 2(Ω · r)ri
dΩi
Z
∂T
= Ωj ρ(r)(δij r2 − rj ri )dV = Iij Ωj
∂Ωi
If we decide to adopt a variable whose time derivative is Ω (that would be the rotation
matrices) then our canonical momenta are the angular momentum vectors pi = Iij Ωj and
1
H = pi I(q)−1
ij pj
2
The associated angles are those of the rotations that are hidden in the moment of inertia
matrix.
Perhaps we can find canonical momenta associated with time derivatives of the quater-
nion. Let q describe body orientation and q = q0 + Q with Q the vector part of Q and
q̇ = q̇0 + Q̇. Because the quaturnion must remain a unitpquaternion we can consider only
Q as specifying the entire quaturnion, with q0 (Q) = 1 − Q2 . We perturb q with an
infinitesimal time dependent rotation specified by Ω. Using quaternion multiplication
d 1
q̇ = [(1 + Ωt/2)q] = [−(Ω · Q) + q0 Ω + Ω × Q] (136)
dt 2
is not the same as Ω. The vector part
q0 qz −qy Ωx
1 1
Q̇ = [q0 Ω + Ω × Q] = −qz q0 qx Ω y
2 2
qy −qx q0 Ωz
By inverting this I could write the Hamiltonian in terms of Q̇. Then I can take derivatives
of the Hamiltonian and find a set of canonical momenta that are conjugate to the vector
part of the quaternion. I invert this and find
2
(q0 + qx2 )q̇x + (−q0 qz + qx qy )q̇y + (q0 qy + qx qz )q̇z
Ω(q, q̇) = 2q0−1 (q0 qz + qx qy )q̇x + (q02 + qy2 )q̇y + (−q0 qx + qy qz )q̇z
(qx qz − q0 qy )q̇x + (q0 qx + qy qz )q̇y + (q02 + qz2 )q̇z
55
∂Ωi
= 2q0−1 [q02 δij − q0 ijk qk + qi qj ]
∂ q̇j
The kinetic energy
1
T (Q, Q̇) = ΩIΩ
2
and Ω(Q̇, Q) and I(Q).
∂T
= IΩ = L
∂Ω
in terms of the angular momentum with I(Q) and Ω(Q, Q̇).
∂T ∂T ∂Ωi ∂Ωi
= = Iil Ωl (139)
∂ q̇j ∂Ωi ∂ q̇j ∂ q̇j
πj = 2(Lj q0 + ijk Li qk + Li qi qj /q0 ) (140)
∂T
= π = 2(Lq0 + L × Q + (L · Q)Qq0−1 ) (141)
∂ Q̇
With T = 21 LI−1 L we can write T (π, Q) and giving canonical coordinates based on
the quaturnion vector Q. However, this Hamiltonian has no obvious conserved quantities,
and we know that angular momentum should be conserved.
In summary it is possible to use Q the vector part of the quaturnion as a canonical
variable and write out Lagrange’s equations in terms of Q, Q̇. It is also possible to derive
a canonical momentum π that is conjugate to Q and an associated Hamiltonian that is
the kinetic energy that is a function of T = H(π, Q). It has the advantage that we
make no assumptions about preferred directions for the angles. However it lacks in obvious
conserved quantities and seems messy, and this probably accounts for its lack of popularity.
Also my choice of Q as variable gives us a potential problem when q0 = 0. This happens
56
when there is a rotation of π/2 about a principal axes. One possible way to redo this might
be to work with 4 degrees of freedom with the quaternion and treat its unit length as a
constraint. Constraints are sometimes ugly to work with in the Hamiltonian view, as you
need to work on a submanifold in phase space.
3.13 A Lie-Poisson numerical method for Solid Body rotation with Quaturnions
We keep track of the body orientation with a quaturnion q(t). The body’s spin vector is
Ω(t). Chose a timestep dt for updating q and Ω(t). In the absence of extra torques the
spin angular momentum vector p is conserved.
Start with orientation defined with quaturnion q(t = 0). The quaturnion should let
you rotate the body from a coordinate with aligned principal axes to the orientation of the
body in an inertial frame. Use initial conditions (current orientation and spin vector) to
compute the initial angular momentum vector p.
• Use the current quaturnion q(t) to compute the inverse of the moment of inertia
matrix I−1 (q) from the quaternion rotation. You don’t need to integrate over the
mass of the body, only rotate or transfer the moment of inertia matrix (tensor) from
the body frame into the current orientation frame.
• Compute the new quaternion q(t + dt) = qdt q(t) via quaturnion multiplication. This
updates q, describing body orientation at the next timestep.
57
requires relating the Andoyer-Deprit angles to ordinary coordinates, and this involves a
series of rotations that depends on the current body orientation and the angular momen-
tum vector with respect to an inertial coordinate system. Methods using Euler angles or
Andoyer-Deprit variables suffer from potential gimbal lock and a quaternion method would
not.
A good reference is the 1994 paper by Jihad Touma and J. Wisdom on ‘Lie-Poisson
integrators for Rigid Body dynamics in the Solar System’, AJ 107, 1189. Possibly Sussman
and Wisdom’s book discuss a similar approach to making a quaternion integrator. (Not
sure as I haven’t found the book, but I saw a reference to it).
4 Tidal Evolution
4.1 Gravitational Potential Expansion in Legendre Polynomials
External to a body where the mass density is zero, the gravitational potential satisfies
Laplace’s equation
∇2 Φ = 0
Working in spherical coordinates r, θ, φ, Laplace’s equation becomes
1 ∂2Φ
∂ 2 ∂Φ ∂ 2 ∂Φ
r + (1 − µ ) + =0
∂r ∂r ∂µ ∂µ 1 − µ2 ∂φ2
where
µ ≡ cos θ
We assume a trial solution
Vn = rn Sn (µ, φ)
The radial derivative term of the trial solution is this
∂ ∂Vn
r2 = n(n + 1)rn Sn
∂r ∂r
Subbing this into Laplace’s equation gives
∂ 2 ∂Sn ∂ 2 ∂Φ
(1 − µ ) + (1 − µ ) + n(n + 1)Sn = 0
∂µ ∂µ ∂µ ∂µ
This equation is known as Legendre’s equation. The functions that satisfy it, Sn , are known
as spherical harmonics. We can expand the general solution in spherical harmonics
X
Φ(r, φ, θ) = An rn + Bn r−(n+1) Sn (µ, θ)
n
58
If the problem has axial symmetry then we can ignore the dependence on φ. In this
case Sn (µ, φ) = Pn (µ) is a Legendre polynomial satisfying
∂ 2 Pn ∂Pn
(1 − µ2 ) 2
− 2µ + n(n + 1)Pn = 0
∂µ ∂µ
Legendre polynomials can be computed with Rodrigues’s formula
1 dn (µ2 − 1)n
Pn (µ) =
2n n! dµn
and the first few are
P0 (µ) = 1
P1 (µ) = µ = cos θ
1 1
P2 (µ) = (3µ2 − 1) = (3 cos 2θ + 1)
2 4
1 3
P3 (µ) = (5µ − 3µ)
2
1
P4 (µ) = (35µ4 − 30µ2 + 3)
8
The gravitational potential of an axisymmetric body with mass M and radius R and
at r > R (external to the body) is often written as
∞
" n #
GM X R
Φ(r, θ) = − 1− Jn Pn (cos θ) (142)
r r
n=2
with coefficients Jn dependent on the internal mass distribution and body shape. Oblate
bodies have positive J2 . A comparison of this equation with equation 114 implies that
(Ik − I⊥ )
J2 =
M R2
(and I think the sign is correct).
59
p p
|r − r0 | = r2 + r02 − 2r · r0 = r2 + r02 − 2rr0 cos µ
with
r · r0
cos µ ≡
rr0
The potential has axial symmetry. There is an important direction, that connecting the
origin and r0 that is the axis of symmetry.
Using the Legendre polynomials
∞
X rl
Φ(r) = Φ(r, µ) = −GMp Pl (µ)
r0l+1
l=0
GMp r2
VT (r, µ) = P2 (µ)
D3
We assume the surface of the deformed body is an equipotential surface.
gh(µ) ∼ VT (R, µ)
where g = GM/R2 is the surface acceleration and h(µ) is the surface height. We can solve
for the surface height giving
h(µ) Mp R 3
= P2 (µ)
R M D3
60
4.4 Love numbers for fluid bodies
A homogenous density nearly spherical body of mean radius C has surface radius
The coefficients for both expressions can be called Love numbers. They are related by
3
kl = hl (147)
2l + 1
They depend on l which is the multipole index. Above I have only included l = 2, 3 terms
but higher order expansions should be consistent with this expression. This expression
is derived by expanding 1/∆ where ∆ is the distance between a point in the shell and a
distant point θ, φ and then integrating over the mass shell. The integration is easier using
spherical harmonics. This is done by M+D with unclearly defined (vaguely normalized)
spherical harmonics.
To summarize: With a deformed body described by the Love numbers for the surface
deformation, hl , we can find the external gravitational potential of the deformed body and
it is described with Love numbers kl .
The tidal potential from an external point mass body of mass mpert expanded in the
coordinate system at the center of M is
Gmpert r 2 r 3
V (r, θ, φ) = − P2 (cos θ) + P3 (cos θ) (148)
a a a
where the tidal perturber is oriented along θ = 0 and is a distance a away from M . The
surface of a tidally deformed fluid body should be an equipotential surface, so the two
expressions (equation146 and 148) for the potential should be equal on the surface.
On the surface (where r = R(θ, φ)), equation 146 gives
GM 3 3
V (θ, φ) = − 1 − h2 P2 (cos θ) − h3 P3 (cos θ) + h2 P2 (cos θ) + h3 P3 (cos θ) (149)
C 5 7
GM 2 4
=− 1 + h2 P2 (cos θ) + h3 P3 (cos θ) (150)
C 5 7
61
where I have expanded to first order in the Love numbers. The coefficients are
3 2l − 2
1− = (151)
2l + 1 2l + 1
On the surface equation 148 gives
" 2 3 #
Gmpert C C
V (θ, φ) = − P2 (cos θ) + P3 (cos θ) . (152)
a a a
Here we just set r = C because mpert is small so we only take the zero-th order term. We
match coefficients and find Love numbers
5 mpert C 3
h2 = (153)
2 M a
7 mpert C 4
h3 = (154)
4 M a
The coefficients are l+1
2l + 1 m C
hl = (155)
2l − 2 M a
Inserting these back into equation 146 gives
l+1
3 mpert C
kl = (156)
2(l + 1) M a
and this agrees with expressions by Efroimsky+12, though the factors of mpert /M and C/a
are not included in their coefficients.
What is the external gravitational potential of our tidally perturbed fluid body?
!
3 mpert C 3 C 2 3 mpert C 4 C 3
GM
V (r, θ, φ) = − 1+ P2 (cos θ) + P3 (cos θ)
r 5 M a r 7 M a r
(157)
The dependence on distance and mass ratio is usually not included in the Love numbers
so that they are independent of the tidal perturbation. If we follow this convention and
evaluate the potential at r = a,
!
GM X mpert C 2l+1
V (a, θ, φ) = − 1+ kl Pl (cos θ) (158)
a M a
l
This is the potential from M at the location of mpert due to tidal deformation of M by the
tide caused by mpert . The quadrupolar term is the strongest one
GM mpert C 5
V2 ∼ −k2 P2 (cos θ) (159)
a M a
We will use this below when we estimate tidally generated torque on a spinning body.
62
4.4.1 Love numbers for an incompressible homogeneous elastic body
When the body is not a fluid, the deformation is lower (h2 is lower) and the Love number
h2 is inversely proportional to the rigidity or the elastic modulus of the deformed body.
Stiffer bodies deform less than stretchy ones. For an incompressible homogeneous elastic
body the Love number
k2 ∼ 0.038eg /µ
where µ is the rigidity and
GM 2
eg =
R4
and R = C. The other Love number h2 can be computed using equation 147.
63
(equation 142) where µ = cos θ is the colatitude and R is the equatorial radius. There is
no n = 1 term as we set the coordinate system to have origin at the center of mass. For
an axisymmetric body
C −A
J2 =
M R2
where C, A are body principal moments of inertia with C > A. For an oblate body, C
corresponds to that for the short axis which is also the axis of symmetry. The coefficient
Z R Z 1
1 2+n
Jn = 2πr dr dµ Pn (µ)ρ(r, µ)
M Rn 0 −1
GM R2 1 2 2
GM
V (r, µ) = − + + Ω r P2 (µ).
r r3 3
where is small. We expand the potential on the surface and assume that it is an equipo-
tential and ignore terms that are not ∝ P2 ,
GM GM GM 1 2 2
V (rs (µ)) = − + P2 (µ) + J2 + Ω R P2 (µ) = constant. (161)
R R R 3
We solve for
1 Ω2 R 3 q
− = J2 + = J2 + (162)
3 GM 3
where
Ω2 R 3
q≡
GM
64
is a measure of centrifugal support. If q = 1, the body flies apart.
Using the formula for P2 (µ) = 12 (3µ2 − 1) we compute the polar and equatorial radii,
the surface has at the pole with θ = 0, µ = 1, rs,polar = R(1 + ). At the equator, θ = π/2,
µ = 0, rs,equatorial = R(1 − /2). A measure of the oblateness
requatorial − rpolar 3
f≡ = − .
requatorial 2
Figure 13: Illustration of the lag angle. The body on the left tidally perturbs the one on
the right which becomes elongated. But since it is spinning its deformation is not exactly
aligned with the perturber. This gives a torque on the spinning body causing it to spin
down or up. Because angular momentum is conserved, this also caused the semi-major
axis a to drift.
We assume principal axis rotation of a spherical homogeneous moon with mass M that
is perturbed tidally by a body M∗ . We assume that they are in circular orbit with semi-
major axis a and zero obliquity. The radius of M is R. We assume that M is not tidally
locked (is not in a spin synchronous state).
We estimate the torque on M∗ by the quadrupolar potential term from M ’s deformation
assuming there is a slight angular shift in V2 given by equation 159 (but with mpert = M∗
65
and C = R). The potential term (equation 159)
5
GM∗ R
V2 ∼ −k2 P2 (cos θ) (164)
a a
The secular part of the semi-diurnal (l = 2) term in the Fourier expansion of the
perturbing potential from point mass M∗ , gives a tidally induced torque
5
3 GM∗2 R
T = k2 sin 2
2 a a
where k2 is the Love number and k2 sin 2 is known as the quality function. The quality
function is often approximated as k2 /Q where Q is a dissipation factor and k2 is a Love
number (but with out factors that depend on mass ratio or distance).
Because the torque is equal and opposite, this torque is exerted on the spinning body
M causing its spin down.
T ∼ I ω̇
where I is the moment of inertia about the spin axis and ω is the spin rate.
So far we have not made any approximations that depend upon whether M is larger or
smaller than M∗ . With M∗ > M the orbital period is P = 2π/n where mean motion
p
n = GM∗ /a3 .
66
Spin down times are often written in terms of P .
ω0
tdespin ≈
ω̇
where ω0 is the initial spin rate. We can approximate our spinning body as a sphere with
moment of inertia I ∼ 25 M R2 . If we chose ω0 = GM/R3 equal to a centrifugal break up
p
GM 52 M R2
p r
ω0 GM/R3 I
tdespin ≈ = =
ω̇ I ω̇ R3 T
3 9
P M 2 a 2 Q
=
15π M∗ R k2
Recent estimates for shear modulus and dissipation factor for icy or rocky rubble give
µQ ∼ 1011 Pa (e.g., Pravic et al. 2014). This is consistent with shear modulus µ = 1 GPa
and dissipation factor Q = 100.
Because angular momentum is conserved there is a relation between the spin down
rate and the drift rate
√ in orbital semi-major axis. If M∗ > M then the orbital angular
momentum L ≈ M GM∗ a and
r
M GM∗ M nȧa
T = L̇ = ȧ = .
2 a 2
With this we can estimate the drift rate in orbital semi-major axis ȧ from the torque
associated with spin down (in the case that the spinning object M < M∗ is low mass)
a M na2
tdrif t = =
ȧ 2T
1 M a 5 Q
=
3n M∗ R k2
P M a 5 Q
=
6π M∗ R k2
67
4.6.2 Associated orbital drift rate during spin down
Using conservation of angular momentum
5
3k2 M R
ȧ = sign(ω − n) na
Q M∗ a
ωlib ∼ αn
5 Spin Resonances
Two approaches to spin resonance. You can expand the potential perturbation taking into
account the orientation of the body. Then with some work you force a Hamiltonian to look
like a pendulum model. Alternatively you compute the torque on the body and try to find
the equations of motion for the body spin/orientation. You also try to force the equations
of motion to look like a pendulum model.
68
5.1 Types of Spin Resonance
A number of possible types of resonance are possible. I will list them!
2. Spin-secular resonance. There is a relation between the planet’s spin precession rate
and an orbital precession rate (like Ω̇) induced by other planets. Saturn’s obliquity
is explained with such a model by Bill Ward and Doug Hamilton where the obliquity
is increased by drifting within such a resonance.
3. Spin-Spin resonance. Two non-spherical asteroids in orbit about each other. Their
spins are multiples of each other.
4. Spin-Binary resonance. Pluto-Charon’s binary orbital frequency could affect the spin
states of Pluto and Charon’s minor satellites.
j
(n − nb ) = w − nb
2
See recent work by Alexandre Correia.
5. Mean motion-spin precession. A mean motion resonance between two objects some-
how affects the spin of one of them. We saw this in some of our simulations for Pluto
and Charon’s minor satellites (and we wrote in 2017 about it).
6. Wobble excitation resonance. The same simulations also showed Pluto and Charon’s
binary perturbations exciting wobbling primarily in the minor satellite Kerberos (but
also sometimes in Styx).
69
Figure 14: Angles relevant for spin orbit resonance
70
summing over non-zero half integer p and with n the mean motion. The coefficients H(p, e)
are polynomials in eccentricity e. To lowest order in e
H(1, e) = 1
H(1/2, e) = −e/2
H(3/2, e) = 7e/2
H(2, e) = 17e2 /2
The higher the half integer |p|, the higher the power of eccentricity in the lowest term;
H(p, e) = O(e2|p−1| )
It is convenient to transfer to a new angle γ = θ − pM for a single half integer p. Then
the equation of motion (and taking only a single sine term) becomes
3 (B − A) hT i
γ̈ + n2 H(p, e) sin 2γ =
2 C C
Resonance capture can be studied using tidal evolution for the averaged torque hT i. Re-
cent work showing that all terms need to be taken into account to estimate the capture
probability which is strongly dependent on body shape (through the moments of inertia)
and eccentricity.
The resonances are separated by half integers of mean motion. q The spin synchronous
or tidal lock term has p = 1 and has libration width dependent on n B−A C . The 3/2 term
q
has libration width that depends n e B−A C . Using widths and distance between resonance
Jack Wisdom developed a resonance overlap criterion for chaotic behavior that was applied
to Hyperion. Apparently once the system is chaotic it is also attitude unstable and so will
start tumbling.
71
6 Problems
1. Reference frames for a symmetric top
Consider a rigid axi-symmetric body with no external forces. In a coordinate frame
aligned with body’s axis of symmetry the moment of inertia matrix has 3 eigenvalues
Ik , I⊥ , I⊥ and two of the them are the same. If the axis of symmetry lies along the z
axis then Ik = Iz . The body is spinning but with angular momentum and spin not
parallel to z or in the xy plane. The body is not spinning about a principal body
axis. Euler’s equations are
a) Show that in the inertial frame that the angular momentum vector is constant.
b) What is the spin vector in the inertial frame?
c) What are the trajectories of the body x, y, z axes as seen in an inertial frame?
Hint:
dL ∂L
= +ω×L
dt ∂t
2. Quadrupole moment and moment of Inertia:
The Quadrupole moment tensor of an extended mass mass distribution with density
ρ(x) is Z
Qαβ ≡ ρ(x)(3xα xβ − δαβ r2 ) d3 x
72
where r2 = 2
P
i xi . The moment of inertia tensor is
Z
Iαβ ≡ ρ(x)(δαβ r2 − xα xβ ) d3 x.
Consider a point mass of mass M at the origin. And consider an extended body of
mass m at coordinate position x = R. The gravitational potential of m is approxi-
mately
Gm GQαβ Rα Rβ
Φ(R) = − −
R 2R5
using summation notation (every pair of indices is summed over each coordinate) and
only expanding to the quadrupolar term.
Show that the gravitational force on point mass M from extended mass m is F with
components Fi
Fi GmRi GQαi Rα 5 Ri
= − + GQαβ Rα Rβ
M R3 R5 2 R7
Show that the torque on M has components
GRj Qkα Rα
Ti = ijk Rj Fk = −M ijk
R5
where ijk the Levi-Civita symbol.
What is the torque on m?
Show that we can replace Q with I giving
Rj GIkα Rα
Ti = 3M ijk
R5
(hint: use symmetry or antisymmetry to drop terms containing the trace).
Show that this can be derived from MacCullagh’s formula.
73
rate and k2 is a Love number. The timescale does not depend upon an external tidal
field. This formula I think is from Peale 1977 in a book called Planetary Satellites
that I cannot get an electronic version of, though it might be in the library.
A similar formula is by Burns and Safronov (1973) (Burns, J.A., and Safronov, V.S.
1973. Asteroid nutation angles. Monthly Notices of the Royal Astronomical Society
of London, Vol. 165, pp. 403 - 411). and is
µQ
τ∼ A
ρR2 w3
with constant A ∼ 100. Here µ, Q, ρ, R, w are the mean shear rigidity, quality factor,
density, body radius and spin rate.
The two formulas are consistent with each other if we use C ∼ mR2 and k2 ∼
Gm2 R−4 µ−1 .
Explain the form of this timescale to order of magnitude.
Hints: Estimate variations in stress, σ, and strain on the body. Recall that stress is
strain times elastic modulus σ ∼ µ. Energy dissipation per unit volume of an elastic
material ė ∼ σ .
˙
5. Averaging a binary
Consider a rapidly spinning binary comprised of masses m1 , m2 separated by distance
a.
Compute the moment of inertia matrix of a ring with mass m and radius r about its
center of mass.
74
If our binary is rapidly spinning (rotation in a plane) while maintaining the separation
a, what is the average value of the moment of inertia matrix (about the center of
mass)? In other words what is
Z P
1
hIi = dt I
P 0
6. Spin Precession
Consider a non-round body of mass m spinning about its maximum principal axis,
the direction of its maximum moment of inertia, equivalently the body’s minimum
axis.
7. Tidal evolution
For a two-body system of mass M and m the total orbital angular momentum is
p
Lo = µ G(M + m)a(1 − e2 )
where a and e are the semi-major axis and eccentricity and µ = mM/(m + M ) is the
reduced mass. Here L is not angular momentum per unit mass. The total orbital
energy is
G(M + m)
Eo = −µ
2a
75
With spin parallel to orbital angular momentum and principal axis rotation, the spin
of m has angular momentum
Lsm = Cm ωm
where Cm is the moment of inertia about the spin axis for mass m and ωm is m’s
spin. The kinetic energy associated with m’s spin is
Cm 2
Esm = ω
2 m
We can describe the spin of body M with moment of inertia CM and spin ωM .
Body deformation associated with tides dissipate energy in m so total energy (orbit
+ spin) is not conserved however the total angular momentum (orbit + spin) is
conserved.
(a) Consider m only dissipating energy (let M be a point mass) and ωm 6= n where
n is the mean motion. When ωm = n the system is tidally locked or in a
spin-synchronous state. What sign is ȧ? Is ė important?
Hint: consider the sign of ωm − n? Is m speeding up or slowing down?
(b) Consider m only dissipating and ωm = n but e > 0. What sign is ȧ? Is ė
important? If so what sign is ė?
(c) Io is drifting outwards (ȧ > 0) but is tidally locked with Jupiter. Why? What
mechanism maintains Io’s non-zero eccentricity? What type of tide is responsible
for Io’s vulcanism? Jupiter is spinning and Io is in an eccentric orbit.
(d) The Moon is drifting away from the Earth. Which type of tide is responsible?
76
9. Excitation of Tumbling
Consider an axi-symmetric torque free body initially spinning about a principal body
axis. Recall (equations 67, 69) with precession rate
#1
Ik2
" !
2
2
φ̇ = ω 1 + 2 − 1 cos α (180)
I⊥
where α is the angle between spin vector and body principal axis. Since we are
considering spinning about a principal body axis consider α small. In the limit of
small α and small θ (angle between angular momentum vector and body axis)
φ̇ Ik
≈
ω I⊥
Ik
ψ̇
≈ 1−
ω I⊥
Suppose you can perturb the body with an oscillating torque (a torque that averages
to zero). What frequency would excite tumbling?
77
What accretion rate would keep Pan spinning and no longer in tidal lock with Saturn?
How fast does Pan’s moment of inertia matrix change due to accretion?
78