You are on page 1of 14

R

Mathematics of
Finite Rotations

R–1
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–2

This Appendix provides a compendium of formulas and results the mathematical treatment of finite
rotations in 3D space. Emphasis is placed on matrix representations of use in the corotational
formulation of finite elements.

§R.1 PLANE VS. SPATIAL ROTATIONS


Plane rotations make up an easy topic. A rotation in, say the x y plane, is defined by just a scalar: the
rotation angle θ about z. Plane rotations commute: θ1 + θ2 = θ2 + θ1 , because the θs are numbers.
The study of spatial rotations is more difficult. The subject is dominated by the fundamental theorem
of Euler:

The general displacement of a rigid body with one point fixed is a rotation
about some axis which passes through that point.

Consequently 3D rotations have both magnitude: the angle of rotation, and direction: the axis of
rotation. These are nominally the same two attributes that categorize vectors. Not coincidentally,
rotations are often depicted as vectors but with a double arrow.
Finite spatial rotations, however, do not obey the laws of vector calculus, although infinitesimal
rotations do. Most striking is failure of commutativity: switching two successive rotations does
not yield the same answer unless the axis of rotation is kept fixed. Within the framework of matrix
algebra, finite rotations can be represented as either 3 × 3 real orthogonal matrices R called rotators
or as real 3 × 3 skew-symmetric matrices Ω called spinors.
The spinor representation is important in physical modeling and theoretical derivations, because
the matrix entries are closely related to the two foregoing attributes. The rotator representation is
important in numerical computations, as well as being naturally related to polar factorizations of a
transformation matrix. The two representations are related by various transformations illustrated
in Figure R.1. Of these the Cayley transform (1857) is the oldest although not the most important
one.

Not unique: adjustable


Unique
by a scale factor γ

Spin(ω) Rot(Ω)
Axial-vector or
Spinor Rotator
pseudo-vector
ω Ω R
axial(Ω) Skew(R)

Figure R.1. Representations of finite space rotations and mapping operations.

Now a 3 × 3 skew-symmetric matrix is defined by three scalar parameters. These three numbers
can be arranged as components of an axial vector ω. Although ω looks like a vector,it does
not obey certain properties of classical vectors such as the composition rule. Therefore the term
pseudo-vector is sometimes used for ω.

R–2
R–3 §R.2 SPINORS

Rotation axis defined


(a) by axial vector
C Can be normalized
in various ways (b)

r Q(x,θ)
θ 2r cos θ/2
P(x)
θ/2 2r sin θ/2
x Q(x,θ) C
x3
xθ θ/2
r
O P(x)
x1 x2

Figure R.2. Attributes of rotation in 3D.

This overview is intended to stress that finite 3D rotations can appear in various mathematical
representations, as depicted in Figure R.1. The exposition that follows expands on this topic, and
studies the connector links shown in Figure R.1.

§R.2 SPINORS
Figure R.2(a) depicts a 3D rotation in space (x1 , x2 , x3 ) by an angle θ about an axis of rotation ω.

For convenience the origin of coordinates O is placed on ω.  The rotation axis is defined by three
directors: ω1 , ω2 , ω3 , at least one of which must be nonzero. These numbers may be scaled by an
nonzero factor γ through which the vector may be normalized in various ways as discussed later.
The positive sense of θ obeys the RHS screw rule.
The rotation takes an arbitrary point P(x), located by its position vector x, into Q(x, θ), located by
its position vector xθ . The center of rotation C is defined by projecting P on the rotation axis. The
plane of rotation CPQ is normal to that axis at C. The radius of rotation is vector r of magnitude r
from C to P. As illustrated in Figure R.2(b) the distance between P and Q is 2r sin 12 θ.

§R.2.1 Spin Matrix and Axial Vector


Given the three directors ω1 , ω2 , and ω3 of the axis ω, we can associate with it a 3×3 skew-symmetric
matrix Ω, called a spin tensor or spin matrix, by the rule
 
0 −ω3 ω2
Ω = Spin (ω) =  ω3 0 −ω1  = −ΩT . (R.1)
−ω2 ω3 0

Premultiplication of a vector v by Ω is equivalent to the cross product of ω and v:

=ω
θ  × v ⇒ θ = Ωv. (R.2)

In particular Ωω = 0, as may be directly verified.

R–3
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–4

The converse operation to (R.1) extracts the 3-vector ω, called or pseudovector, or axial vector
from a given spin tensor:  
ω1
ω = axial (Ω) =  ω2  (R.3)
ω3
The length of this vector is denoted by ω:

ω = |ω| = + ω12 + ω22 + ω32 . (R.4)

As general notational rule, we will use corresponding upper and lower case symbols for the spin
matrix and its axial vector, respectively. For example, N and n, Θ and θ, B and b.

§R.2.2 Normalizations
As noted, ω and Ω can be multiplied by a nonzero scalar factor γ to obtain various normalizations.
In general γ has the form g(θ)/ω, where g(.) is a function of the rotation angle θ. The purpose of
normalizations is to simplify the connections Rot and Skew to the rotator, to avoid singularities
for special angles, and to connect the components ω1 , ω2 and ω3 closely to the rotation amplitude.
This section review some normalizations that have practical or historical importance.
Taking γ = 1/ω we obtain the unit axial-vector and unit spinor, which are denoted by n and N,
respectively:
     
n1 ω1 /ω 0 −n 3 n 2
ω Ω
n =  n 2  =  ω2 /ω  = , N = Spin (n) = =  n3 0 −n 1  . (R.5)
ω ω
n3 ω3 /ω −n 2 n 3 0

Taking γ = tan 12 θ/ω is equivalent to multiplying the n i by tan 12 θ. We thus obtain the Rodrigues
parameters bi = n i tan 12 θ, i = 1, 2, 3. These are collected in the Rodrigues axial-vector b with
associated spinor B:

b = tan 12 θ n = (tan 12 θ/ω) ω, B = Spin (b) = tan 12 θ N = (tan 12 θ/ω) Ω. (R.6)

This representation permits an elegant formulation of the rotator via the Cayley transform studied
later. However it collapses as θ nears 180◦ since tan 12 θ → ±∞ as θ ↔ 180◦ . One way to
circumvent the singularity is through the use of the four Euler-Rodrigues parameters, also called
quaternion coefficients:

p0 = cos 12 θ, pi = n i sin 12 θ = ωi /ω sin 12 θ, i = 1, 2, 3. (R.7)

under the constraint p02 + p12 + p22 + p32 = 1. This set is often used in multibody dynamics, robotics
and control. It comes at the cost of carrying along an extra parameter and an additional constraint.
A related singularity-free normalization introduced by DeVeubeke [ ] takes γ = sin 12 θ/ω and is
equivalent to using only the last three parameters of (R.7):

p = sin 12 θ n = (sin 12 θ/ω) ω, P = Spin (p) = sin 12 θ N = (sin 12 θ/ω) Ω. (R.8)

R–4
R–5 §R.2 SPINORS

Finally, an important normalization that preserves three parameters while avoiding singularities is
that associated with the exponential map. Introduce a rotation vector θ defined as

θ = θ n = (θ/ω) ω, Θ = Spin (θ) = θ N = (θ/ω) Ω. (R.9)



For this normalization the angle is the length of the rotation vector: θ = |θ| = θ12 + θ22 + θ32 .
The selection of the sign is a matter of convention.
§R.2.3 Spectral Properties
Study of the spinor eigensystem Ωvi = λi vi is of interest for various developments. Begin by
forming the characteristic equation

det(Ω − λI) = −λ3 − ω2 λ = 0 (R.10)

where I denotes the identity matrix of order 3. It follows that the eigenvalues of Ω are λ1 = 0,
λ2,3 = ±ωi. Consequently Ω is singular with rank 2 if ω = 0 whereas if ω = 0, Ω is null.
The eigenvalues are collected in the diagonal matrix Λ = diag(0, ωi, −ωi) and the corresponding
right eigenvectors vi in columns of V = [ v1 v2 v3 ], so that ΩV = VΛ. A cyclic-symmetric
expression of V, obtained through Mathematica, is
 
ω1 ω1 s − ω2 + i(ω2 − ω3 )ω ω1 s − ω2 − i(ω2 − ω3 )ω

V = ω2 ω2 s − ω2 + i(ω3 − ω1 )ω ω2 s − ω2 − i(ω3 − ω1 )ω  (R.11)
ω3 ω3 s − ω2 + i(ω1 − ω2 )ω ω3 s − ω2 − i(ω1 − ω2 )ω

where s = ω1 + ω2 + ω3 . Its inverse is


 
ω22 + ω32 ω22 + ω32
ω1 − 12 − 12
 ω1 s − ω2 − i(ω2 − ω3 )ω ω1 s − ω2 + i(ω2 − ω3 )ω 
 
1  ω32 + ω12 ω32 + ω12 
V−1 = 2  ω2 − 12 − 12  (R.12)
ω  ω2 s − ω2 − i(ω3 − ω1 )ω ω2 s − ω2 + i(ω3 − ω1 )ω 
 
ω12 + ω22 ω12 + ω22
ω3 − 12 − 12
ω3 s − ω2 − i(ω1 − ω2 )ω ω3 s − ω2 + i(ω1 − ω2 )ω
The real and imaginary part of the eigenvectors v2 and v3 are orthogonal. This is a general property
of skew-symmetric matrices; cf. Bellman [ , p. 64].
Because the eigenvalues of Ω are distinct if ω = 0, an arbitrary matrix function F(Ω) can be
explicitly obtained as
 
f (0) 0 0
F(Ω) = V  0 f (ωi) 0  V−1 , (R.13)
0 0 f (−ωi)

in which f (.) is the scalar version of F(.). One important application of (R.13) is the matrix
exponential: f (.) → e(.) .

R–5
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–6

The square of Ω, computed through direct multiplication, is


 2 
ω2 + ω32 −ω1 ω2 −ω1 ω3
Ω2 = −  −ω1 ω2 ω32 + ω12 −ω2 ω3  = ωωT − ω2 I = ω2 (nnT − I). (R.14)
−ω1 ω3 −ω2 ω3 ω12 + ω22

This is a symmetric matrix of trace −2ω2 whose eigenvalues are 0, −ω2 and −ω2 .
By the Cayley-Hamilton theorem, Ω satisfies its own characteristic equation (R.10):

Ω3 = −ω2 Ω, Ω4 = −ω2 Ω2 , . . . and generally Ωn = −ω2 Ωn−2 , n ≥ 3. (R.15)

Hence if n = 3, 5, . . . the odd powers Ωn are skew-symmetric with distinct purely imaginary eigen-
values, whereas if n = 4, 6 . . ., the even powers Ωn are symmetric with repeated real eigenvalues.
The eigenvalues of I + γ Ω and I − γ Ω, are (1, 1 ± γ ωi) and (−1, 1 ± γ ωi), respectively. Hence
those two matrices are guaranteed to be nonsingular. This has implications in the Cayley transform.

EXAMPLE R.1

Consider the pseudo-vector ω = [ 6 2 3 ]T , for which ω = 62 + 22 + 32 = 7. The associated spin matrix
and its square are

0 −3 2 13 −12 −18
Ω= 3 0 −6 , Ω2 = − −12 45 −6 . (R.16)
−2 6 0 −18 −6 40
The eigenvalues of Ω are (0, 7i, −7i) while those of Ω2 are (0, −7, −7).

§R.3 FROM SPINORS TO ROTATORS


Referring to Figure R.2, a rotator is an operator that maps a generic point P(x) to Q(xθ ) given the
 and the angle θ. We consider only rotator representations in the form of rotation
rotation axis ω
matrices R, defined by
xθ = Rx (R.17)
This 3 × 3 matrix is proper orthogonal, that is, RT R = I and det(R) = +1. It must reduce to I if
the rotation vanishes. Another important attribute is the trace property

trace(R) = 1 + 2 cos θ, (R.18)

proofs of which may be found for example in Hammermesh [ , p.325] or Goldstein [ , p. ]. The
problem considered in this section is the construction of R from the rotation data. The inverse
problem: given R, extract ω and θ, is treated in the next section. Now if R is assumed to be analytic
in Ω it must have the Taylor expansion R = I + c1 Ω + c2 Ω2 + c3 Ω3 + . . ., where all ci must vanish
if θ = 0. But in view of the Cayley-Hamilton theorem (R.15), all powers of order 3 or higher may
be eliminated, and so R must be a linear function of I, Ω and Ω2 . For convenience this will be
written
R = I + α(γ Ω) + β(γ Ω)2 , (R.19)

R–6
R–7 §R.3 FROM SPINORS TO ROTATORS

Table R.1. Rotator Forms for Various Spinors

Parametrization γ α β Spinor Rotator R

sin θ 2 sin2 12 θ θ Ω + 2 sin 2 θ Ω2


2 1
None 1 ω Ω I + sin
ω
ω2 ω2
Unit axial-vector 1 sin θ 2 sin2 12 θ N = γΩ I + sin θ N + 2 sin2 12 θ N2 ,
ω
tan 12 θ
Rodrigues-Cayley ω 2 cos2 12 θ 2 cos2 12 θ B = γΩ I + 2 cos2 12 θ (B + B2 ) = (I + B)(I − B)−1
sin 12 θ
DeVeubeke ω 2 cos 12 θ 2 P = γΩ I + 2 cos 12 θ P + 2P2
θ sin θ 2 sin2 12 θ θ Θ + 2 sin 2 θ Θ2 = eΘ = eθ N
2 1
Exponential map ω θ Θ = γ Ω I + sin
θ
θ2 θ2

where γ is the scaling factor discussed above, whereas α and β are scalar functions of θ and of
invariants of Ω or ω. Since the only invariant of the latter is ω we can anticipate that α = α(θ, ω)
and β = β(θ, ω), both vanishing if θ = 0. Two techniques to determine those coefficients for
γ = 1 are discussed next. Table R.1 summarizes the most important representations of the rotator
in terms of the scaled Ω.

§R.3.1 The Algebraic Approach


This approach finds α and β for γ = 1 (the unscaled spinor) directly from algebraic conditions.
Taking the trace of (R.19) and applying the property (R.18) requires

1 − cos θ 2 sin2 21 θ
3 − 2βω = 1 + 2 cos θ,
2
→ β= = . (R.20)
ω2 ω2
The orthogonality condition I = RT R = (I − αΩ + βΩ2 )(I + αΩ + βΩ2 ) = I + (2β − α 2 )Ω2 +
β 2 Ω4 = I + (2β − α 2 − β 2 ω2 )Ω2 leads to
sin θ
2β − α 2 − β 2 ω2 = 0 → α= (R.21)
ω
Hence
sin θ 1 − cos θ 2 sin θ 2 sin2 21 θ 2
R=I+ Ω+ Ω =I+ Ω+ Ω. (R.22)
ω ω2 ω ω2
From a computational viewpoint the sine-squared form should be preferred for small angles to avoid
the 1 − cos θ cancellation. Replacing the components of Ω and Ω2 gives the explicit rotator form

 ω12 + (ω22 + ω32 ) cos θ 2ω1 ω2 sin2 21 θ − ω3 ω sin θ 2ω1 ω3 sin2 12 θ + ω2 ω sin θ 
1
R = 2  2ω1 ω2 sin2 12 θ + ω3 ω sin θ ω22 + (ω32 + ω12 ) cos θ 2ω2 ω3 sin2 12 θ − ω1 ω sin θ  .
ω
2ω1 ω3 sin2 12 θ − ω2 ω sin θ 2ω2 ω3 sin2 21 θ + ω1 ω sin θ ω32 + (ω12 + ω22 ) cos θ
(R.23)
This is invariant to scaling of the ωi , and consequently (R.23) is unique.

R–7
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–8

§R.3.2 The Geometric Approach


The vector representation of the rigid motion depicted in Figure R.2 is

xθ = x cos θ + (
n × x) sin θ + n
 (
n · x)(1 − cos θ) = x + (
n × x) sin θ + n × (
n × x) (1 − cos θ ).
(R.24)
where n is ω
 normalized to unit length as per (R.9). This can be recast in matrix form by substituting
 × x → Nx = Ωx/ω and xθ = Rx. Cancelling x we get back (R.22).
n

§R.3.3 Rotators for Various Parametrizations


If ω is unit-length-normalized to n as per (R.5), γ = 1/ω and R = I + sin θ N + (1 − cos θ ) N2 .
This is the matrix form of (R.23). Because N2 = nnT − I, an ocassionally useful variant is

R = W + (1 − cos θ) n nT , W = cos θ I + sin θ N. (R.25)

In terms of the three Rodrigues-Cayley parameters bi introduced in (R.6), α = β = 2 cos2 12 θ and


R = I + 2 cos2 21 θ (B + B2 ). This can be explicitly worked out to be
 
1 + b12 − b22 − b32 2(b1 b2 − b3 ) 2(b1 b3 + b2 )
1  2(b1 b2 + b3 )
R= 1 − b12 + b22 − b32 2(b2 b3 − b1 )  (R.26)
1 + b1 + b22 + b32
2
2(b1 b3 − b2 ) 2(b2 b3 + b1 ) 1 − b12 − b22 + b32

This form was derived by Rodrigues [ ] and used by Cayley [ ] to study rigid body motions. It has
the advantage of being obtainable through an algebraic matrix expression: the Cayley transform,
presented below. It becomes indeterminate, however, as θ → 180◦ , since all terms approach 0/0.
This indeterminacy is avoided by using the four Euler-Rodrigues parameters, which are also the
quaternion coefficients, defined in (R.6). In terms of these the rotator becomes
 2 
p0 + p12 − 12 p1 p2 − p0 p3 p1 p3 + p0 p2
R = 2  p1 p2 + p0 p3 p02 + p22 − 12 p2 p3 − p0 p1  (R.27)
p1 p3 − p0 p2 p2 p3 + p0 p1 ) p0 + p3 − 2
2 2 1

This rotator cannot become singular, but this is paid at the cost of carrying along an extra parameter
plus the constraint p02 + p12 + p22 + p32 = 1.
The normalization of DeVeubeke (R.8): pi = (ωi /ω) sin 12 θ, leads to α = 2 cos 12 θ and β = 2.
Hence R = I + 2 cos 12 θ P + 2P2 .

§R.3.4 The Cayley Transform


Given any skew-symmetric real matrix S = −ST , we can apply the transformation

Q = (I + S)(I − S)−1 (R.28)

Then Q is a proper orthogonal matrix, that is QT Q = I and det Q = +1. This is stated in
several textbooks, e.g., Gantmacher [ ] but none gives a proof. Here is the proof of orthogonality:

R–8
R–9 §R.3 FROM SPINORS TO ROTATORS

QT Q = (I+S)−1 (I−S)(I+S)(I−S)−1 = (I+S)−1 (I+S)(I−S)(I−S)−1 = II = I because I+S


and I − S commute. The property det Q = +1 can be easily proven from the spectral properties.
The inverse transformation
S = (Q − I)(Q + I)−1 (R.29)

produces skew-symmetric matrices from a source orthogonal matrix Q. Equations (R.28) and
(R.29) are called the Cayley transforms after Cayley [ ]. These formulas are ocassionally useful in
the construction of approximations for moderate rotations.
An interesting question is: given Ω and θ, can (R.28) be used to produce the exact R? The answer is:
yes, if Ω is scaled by a factor γ = γ (θ, ω). We thus investigate whether R = (I + γ Ω)(I − γ Ω)−1
exactly for some γ . Premultiplying both sides by I − γ Ω and representing R by (R.19) we require

(I − γ Ω)(I + αΩ + βΩ2 ) = I + (α − γ + γβω2 )Ω + (β − αγ )Ω2 = I + γ Ω (R.30)

Identifying we get the conditions β = αγ and α − γ + γβω2 = γ . The first one gives γ = β/α,
which inserted in the second requires α 2 − 2β + βω2 = 0. Fortunately this is identically satisfied by
(R.21). Thus the only solution is γ = β/α = (1 − cos θ)/(ω sin θ) = tan 12 θ/ω. This is precisely
the Rodrigues normalization (R.7). Consequently

−1
tan 12 θ
R = (I + B)(I − B) , B= Ω. (R.31)
ω

The explicit calculation of R in terms of the bi leads to (R.26).

§R.3.5 Exponential Map

This is a final representation of R that has both theoretical and practical importance. Given a
skew-symmetric real matrix S, the matrix exponential

Q = eS = Exp(S) (R.32)

is proper orthogonal. Here is the simple proof of Gantmacher [ , p 287]: QT = Exp(ST ) =


−1
Exp(−S ) = Q . If the eigenvalues of S are
T
λi , i λi = trace(S) = 0. The eigenvalues of Q are
µi = exp(λi ); thus det(Q) = i µi = exp( i λi ) = exp(0) = +1. The transformation (R.32) is
called an exponential map. The converse is of course S = Log(Q).
As in the case of the Cayley transform, one may pose the question of whether we can get the rotator
R = Exp(γ Ω) exactly for some factor γ = γ (θ, ω). To study this question we need an explicit
form of the exponential. This can be obtained from (R.13) in which the function Exp so that the
diagonal matrix entries are 1 and exp(±γ ωi) = cos γ ω ±i sin γ ω. The following approach is more
instructive and leads directly to the final result. Start from the definition of the matrix exponential

γ2 2 γ3 3
Exp(γ Ω) = I + γ Ω + Ω + Ω + ... (R.33)
2! 3!
R–9
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–10

and use the Cayley-Hamilton theorem (R.15) to eliminate all powers of order 3 or higher in Ω.
Identify the coefficient series of Ω and Ω2 with those of the sine and cosine, to obtain

sin(γ ω) 1 − cos(γ ω) 2
Exp(γ Ω) = I + Ω+ Ω. (R.34)
ω ω2
Comparing to (R.22) requires γ ω = θ, or γ = θ/ω. Introducing θi = θωi /ω and Θ = Spin (θ) =
θN = (θ/ω)Ω as in (R.9), one gets

sin θ 1 − cos θ 2 sin θ 2 sin2 21 θ 2


R = Exp(Θ) = I + Θ+ Θ = I + Θ + Θ . (R.35)
θ θ2 θ θ2

On substituting Θ = θN this recovers R = I + sin θ N + (1 − cos θ) N2 , as it should.


This representation has several advantages: (i) it is singularity free, and (ii) the θi are exactly
proportional to the angle, and (ii) it simplifies differentiation. Because of these favorable properties
the exponential map has become a favorite of implementations where large angles may occur, as in
orbiting structures and robotics.

EXAMPLE R.2

Take again ω = [ 3 2 6 ]T so ω = 7. Consider the 3 angles θ = 2◦ , θ = 90◦ and θ = 180◦ . The rotators
calculated from (R.22) are, to 8 figures:
R= (R.36)

§R.4 FROM ROTATORS TO SPINORS


If R is given, the extraction of the rotation amount θ and the unit pseudo-vector n = ω/ω is often
required. The former is easy using the trace property (R.19):

cos θ = 1
2 (trace(R) − 1) (R.37)

Recovery of n is also straightforward using the particular form R = I + sin θ N + (1 − cos θ )N2
since R − RT = 2 sin θ N, whence

R − RT
N= , n = axial (N) (R.38)
2 sin θ
One issue is the sign of θ since (R.37) is satisfied by ±θ. If the sign is reversed, so is n. Thus in
principle it is possible to select θ ≥ 0 if no constraints are placed on the direction of the rotation
axis.
The above formulas are prone to numerical instability for θ near 0◦ and 180◦ since sin θ vanishes. A
robust algorithm is that given by Spurrier [ ] in the language of quaternions. Choose the algebraically
largest of trace(R) and Rii , i = 1, 2, 3. If trace(R) is the largest, compute

p0 = cos 12 θ = 12 1 + trace(R), pi = n i sin 12 θ = 14 (Rk j − R jk )/ p0 , i = 1, 2, 3, (R.39)

R–10
R–11 §R.5 ROTATOR DERIVATIVES

where j and k are the cyclic permutations of i. Otherwise let Rii be the algebraically largest
diagonal entry, and again denote i, j, k the cyclic permutation. Then use

pi = n i sin 2 θ = 12 Mii + 14 (1 − trace(M)),
1
p0 = cos 12 θ = 14 (Rk j − R jk )/ pi ,
(R.40)
pi = 14 (Rl,i + Ri,l )/ pi , l = j, k.

From p0 , p1 , p2 , p3 it is easy to pass to θ, n1, n2, n3 once the sign of θ is chosen as discussed
above.
REMARK R.1
The theoretical formula for the matrix logarithm, which applies to any matrix size, is

arcsin τ
Ω = log R = axial (R − RT ), (R.41)

where τ = 12 |axial (R − RT )|. But for numerical computations this expression is largely useless.

§R.5 ROTATOR DERIVATIVES


The derivatives and differentials of the rotator with respect to angle and rotation axis direction
changes are required in many developments.
As independent parameters we will take θ and ω = ωn (the n i are not good choices because they
are linked by a constraint). For convenience, define
   2 
dω1 d ω1
dω = dω2 , dΩ = Spin (dω), d ω = d 2 ω2  , d 2 Ω = Spin (d 2 ω),
  2 
dω3 d 2 ω3 (R.42)

W = cos θ I + sin θ N = R − (1 − cos θ) nnT .

It is convenient to depart from R = Exp(θ N) = I+sin θ N+(1−cos θ) N2 = W+(1−cos θ )nnT .


Succesive differentiation gives

dR = R d(θ N) = R (N dθ + θ dN), (R.43)



d 2 R = R d(θN) d(θ N) + R d 2 (θ N) = R (N dθ + θ dN)2 + N d 2 θ + 2 dθ dN + θ d 2 N
(R.44)
These expressions can be simplified using the following identities:

RN = NR = N + sin θ N2 + (1 − cos θ)N3 = cos θ N + sin θ N2 = WN = NW,


NRN = N2 R = R N2 = − sin θ N + cos θ N2 = R − I − N2 = R − nnT ,
dΩ ωT dω dω ωT − (ωT dω) I N dω ωT − (ωT dω) N
dN = −N N dN = , N2 dN = ,
ω ω2 ω2 ω2
d ωT dω
R dN = R − WN .
ω ω2
(R.45)

R–11
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–12

Here Nn = 0 has been used to eliminate some terms.


Inserting into (R.44) gives
dR = WN dθ + θ R dN (R.46)

This is a compacted form verified through Mathematica.


The expression of d 2 R, which is required for stiffness level and acceleration computations, will be
worked out later.

§R.6 COROTATIONAL KINEMATICS


An important application of finite rotations is what is now called the corotational kinemtic descrip-
tion of finite elements. The approach has roots on a very old idea: the separation of rigid body and
purely deformational motions in solid and continuum mechanics. It originally arose in theories of
small deformations coupled with large rigid motions. The pioneering work of Biot [] in this subject
should be cited.
Important technology applications of these theories are aircraft and spacecraft dynamics on the one
hand, and geological problems on the other. The rigid-plus-deformational decomposition idea for
an entire structure was originally used in the 1960s in the context of the dynamics and control of
orbiting space systems as well as aircraft structures. The motivation was essentially to trace the
mean motion.
This approach was systematized by Fraeijs de Veubeke in an important paper [ ], which essentially
closed the subject as regards a complete structure. The motivation for this approach was clearly
expressed by de Veubeke in the Introduction of that paper, which appeared two months before his
untimely death.
“The formulation of the motion of a flexible body as a continuum through intertial space is unsatisfactory
from several viewpoints. One is usually not interested in the details of this motion but in its main
characteristics such as the motion of the center of mass and, under the assumptions that the deformations
remain small, the history of the average orientation of the body. The last information is of course
essential to pilots, real and artificial, in order to implement guidance corrections. We therefore try
to define a set of Cartesian mean axes accompanying the body, or dynamic reference frame, with
respect to which the relative displacements, velocities or accelerations of material points due to the
deformations are minimum in some global sense. If the body does not deform, any set of axes fixed
into the body is of course a natural dynamic reference frame.”
In this section we ouline De Vebeuke’s method using the notation and mathematical tools presented
in previous sections. We consider a flexible body, such as an aircraft, moving through space. At each
time t we distinguish three spatial configurations illustrated in Figure R.3. In these configurations
G denotes the center of mass whereas P is an arbitrary point.
1. Base Reference Configuration C0 : referred to a RCC {X 1 , X 2 , X 3 }. In dynamics this is an
inertial (acceleration free) system, which is not necessarily fixed to the ground but may move
at a constant velocity fitted to the orbit or trajectory. In statics it is simply a fixed system. Also
called the global system in FEM applications. The body origin G 0 is placed at the center of
mass.

R–12
R–13 §R.6 COROTATIONAL KINEMATICS

2. Dynamic Reference Configuration C: ¯ obtained by a rigid body motion of C0 . Its position


coordinates are denoted by {x̄1 , x̄2 , x̄3 } referred to a system {x̄1 , x̄2 , x̄3 } which is the dynamic
translation of {X 1 , X 2 , X 3 }. This motion can be tought as a translation of the center of mass
G 0 → Ḡ, followed by a rotation about Ḡ.
3. Current Configuration C: this is the (generally deformed) configuration occupied by the body
at time t. Its position coordinates are denoted by {x1 , x2 , x3 }.
It should be noted that only C is an actual configuration. Both reference configurations 0 and C are
virtual in the sense that they are not generally occupied by the body at any instance of time.
The global displacements from C0 to C are denoted by u = x−X. The deformational displacements
from Cd to C are d = x − xd in the base system and d̄ = x̄ − x̄d in the corotated system.
The dynamic configuration C¯ is defined by the translation a of the center of mass and the rotation
R:
X = a + x + d, X̄ = ā + x̄ + d̄ (R.47)

x = Rx̄, x̄ = RT x. (R.48)
Since x and d  and a are observed in the inertial
 is observed in the dynamic frame x̄i whereas X
frame xi , the most useful relation is

X = a + R(x + d) (R.49)

References

[1] J. H. Argyris, An Excursion into Large Rotations, Comp. Meths. Appl. Mech. Engrg., 32, pp.
85-155, 1982.
[2] R. Bellman, Introduction to Matrix Analysis, McGraw-Hill, New York, 1960
[3] M. A. Biot, The Mechanics of Incremental Deformations, McGraw-Hill, New York, 1962.
[4] H. Chung and K. C. Gupta, An historical note on finite rotations, J. Appl. Mech., 56, pp.
139–145, 1989.
[5] B. Fraeijs De Veubeke, The Dynamics of Flexible Bodies, Int. J. Engrg. Sci, 14, pp. 895-913,
1976.
[6] F. R. Gantmacher, The Theory of Matrices, Vol. I, Chelsea, New York, 1960.
[7] A. S. Householder, The Theory of Matrices in Numerical Analysis, Blaisdell, New York, 1964.
[8] B. Nour-Omid and C C. Rankin, Finite Rotation Analysis and Consistent Linearization Using
Projectors, submitted to Comp. Meths. Appl. Mech. Engrg., 1990.
[9] O. Rodrigues, Des lois géometriques qui regissent les déplacements d’un systéme solide
dans l’espace, et de la variation des coordonnées provenant de ces déplacement considerées

R–13
Appendix R: MATHEMATICS OF FINITE ROTATIONS R–14

_
_ P
_ _ C d
x3 G P
dG
_
uG
G C
_ _
x2 u uG
_ _
x1
xG
_
_ xG _ xG
xG x x u

X3, x 3 P0
G0 C0
X
XG
X1,x1 X2 , x2

Figure R.1. The corotational description for a moving structure. The dynamic reference configuration
system is shown way offset for visualization convenience.

indépendent des causes qui peuvent les produire, J. de Mathématiques Pures et Appliquées, 5,
380–400, 1840
[10] M. L. Szwabowicz, Variational Formulation in the Geometrically Nonlinear Thin Elastic Shell
Theory, Int. J. Engrg. Sci., 22, pp. 1161-1175, 1986.
[11] J. C. Simo, A Finite Strain Beam Formulation. Part I: The Three-Dimensional Dynamic
Problem, Comp. Meths. Appl. Mech. Engrg., 49, pp. 55-70, 1985.
[12] R. A. Spurrier, A Comment on Singularity-Free Extraction of a Quaternion from a Direction
Cosine Matrix, J. Spacecrafts & Rockets, 15, p. 255, 1978.
[13] R. Turnbull, The Theory of Determinants, Matrices and Invariants, Blackie and Son, London,
1929.

R–14

You might also like