Movatterモバイル変換


[0]ホーム

URL:


Wikipedia

3D rotation group

Inmechanics andgeometry, the3D rotation group, often denotedSO(3), is thegroup of allrotations about theorigin ofthree-dimensionalEuclidean spaceR3{\displaystyle \mathbb {R} ^{3}} under the operation ofcomposition.[1]

By definition, a rotation about the origin is a transformation that preserves the origin,Euclidean distance (so it is anisometry), andorientation (i.e.,handedness of space). Composing two rotations results in another rotation, every rotation has a uniqueinverse rotation, and theidentity map satisfies the definition of a rotation. Owing to the above properties (along composite rotations'associative property), the set of all rotations is agroup under composition.

Every non-trivial rotation is determined by its axis of rotation (a line through the origin) and its angle of rotation. Rotations are not commutative (for example, rotatingR 90° in the x-y plane followed byS 90° in the y-z plane is not the same asS followed byR), making the 3D rotation group anonabelian group. Moreover, the rotation group has a natural structure as amanifold for which the group operations aresmoothly differentiable, so it is in fact aLie group. It iscompact and has dimension 3.

Rotations arelinear transformations ofR3{\displaystyle \mathbb {R} ^{3}} and can therefore be represented bymatrices once abasis ofR3{\displaystyle \mathbb {R} ^{3}} has been chosen. Specifically, if we choose anorthonormal basis ofR3{\displaystyle \mathbb {R} ^{3}}, every rotation is described by anorthogonal 3 × 3 matrix (i.e., a 3 × 3 matrix with real entries which, when multiplied by itstranspose, results in theidentity matrix) withdeterminant 1. The group SO(3) can therefore be identified with the group of these matrices undermatrix multiplication. These matrices are known as "special orthogonal matrices", explaining the notation SO(3).

The group SO(3) is used to describe the possible rotational symmetries of an object, as well as the possible orientations of an object in space. Itsrepresentations are important in physics, where they give rise to theelementary particles of integerspin.

Length and angle

edit

Besides just preserving length, rotations also preserve theangles between vectors. This follows from the fact that the standarddot product between two vectorsu andv can be written purely in terms of length (see thelaw of cosines):uv=12(u+v2u2v2).{\displaystyle \mathbf {u} \cdot \mathbf {v} ={\frac {1}{2}}\left(\|\mathbf {u} +\mathbf {v} \|^{2}-\|\mathbf {u} \|^{2}-\|\mathbf {v} \|^{2}\right).} 

It follows that every length-preserving linear transformation inR3{\displaystyle \mathbb {R} ^{3}}  preserves the dot product, and thus the angle between vectors. Rotations are often defined as linear transformations that preserve the inner product onR3{\displaystyle \mathbb {R} ^{3}} , which is equivalent to requiring them to preserve length. Seeclassical group for a treatment of this more general approach, whereSO(3) appears as a special case.

Orthogonal and rotation matrices

edit

Every rotation maps anorthonormal basis ofR3{\displaystyle \mathbb {R} ^{3}}  to another orthonormal basis. Like any linear transformation offinite-dimensional vector spaces, a rotation can always be represented by amatrix. LetR be a given rotation. With respect to thestandard basise1,e2,e3 ofR3{\displaystyle \mathbb {R} ^{3}}  the columns ofR are given by(Re1,Re2,Re3). Since the standard basis is orthonormal, and sinceR preserves angles and length, the columns ofR form another orthonormal basis. This orthonormality condition can be expressed in the form

RTR=RRT=I,{\displaystyle R^{\mathsf {T}}R=RR^{\mathsf {T}}=I,} 

whereRT denotes thetranspose ofR andI is the3 × 3identity matrix. Matrices for which this property holds are calledorthogonal matrices. The group of all3 × 3 orthogonal matrices is denotedO(3), and consists of all proper and improper rotations.

In addition to preserving length, proper rotations must also preserve orientation. A matrix will preserve or reverse orientation according to whether thedeterminant of the matrix is positive or negative. For an orthogonal matrixR, note thatdetRT = detR implies(detR)2 = 1, so thatdetR = ±1. Thesubgroup of orthogonal matrices with determinant+1 is called thespecialorthogonal group, denotedSO(3).

Thus every rotation can be represented uniquely by an orthogonal matrix with unit determinant. Moreover, since composition of rotations corresponds tomatrix multiplication, the rotation group isisomorphic to the special orthogonal groupSO(3).

Improper rotations correspond to orthogonal matrices with determinant−1, and they do not form a group because the product of two improper rotations is a proper rotation.

Group structure

edit

The rotation group is agroup underfunction composition (or equivalently theproduct of linear transformations). It is asubgroup of thegeneral linear group consisting of allinvertible linear transformations of thereal 3-spaceR3{\displaystyle \mathbb {R} ^{3}} .[2]

Furthermore, the rotation group isnonabelian. That is, the order in which rotations are composed makes a difference. For example, a quarter turn around the positivex-axis followed by a quarter turn around the positivey-axis is a different rotation than the one obtained by first rotating aroundy and thenx.

The orthogonal group, consisting of all proper and improper rotations, is generated by reflections. Every proper rotation is the composition of two reflections, a special case of theCartan–Dieudonné theorem.

Complete classification of finite subgroups

edit

The finite subgroups ofSO(3){\displaystyle \mathrm {SO} (3)}  are completelyclassified.[3]

Every finite subgroup is isomorphic to either an element of one of twocountably infinite families of planar isometries: thecyclic groupsCn{\displaystyle C_{n}}  or thedihedral groupsD2n{\displaystyle D_{2n}} , or to one of three other groups: thetetrahedral groupA4{\displaystyle \cong A_{4}} , theoctahedral groupS4{\displaystyle \cong S_{4}} , or theicosahedral groupA5{\displaystyle \cong A_{5}} .

Axis of rotation

edit

Every nontrivial proper rotation in 3 dimensions fixes a unique 1-dimensionallinear subspace ofR3{\displaystyle \mathbb {R} ^{3}}  which is called theaxis of rotation (this isEuler's rotation theorem). Each such rotation acts as an ordinary 2-dimensional rotation in the planeorthogonal to this axis. Since every 2-dimensional rotation can be represented by an angleφ, an arbitrary 3-dimensional rotation can be specified by an axis of rotation together with anangle of rotation about this axis. (Technically, one needs to specify an orientation for the axis and whether the rotation is taken to beclockwise orcounterclockwise with respect to this orientation).

For example, counterclockwise rotation about the positivez-axis by angleφ is given by

Rz(ϕ)=[cosϕsinϕ0sinϕcosϕ0001].{\displaystyle R_{z}(\phi )={\begin{bmatrix}\cos \phi &-\sin \phi &0\\\sin \phi &\cos \phi &0\\0&0&1\end{bmatrix}}.} 

Given aunit vectorn inR3{\displaystyle \mathbb {R} ^{3}}  and an angleφ, letR(φ, n) represent a counterclockwise rotation about the axis throughn (with orientation determined byn). Then

  • R(0,n) is the identity transformation for anyn
  • R(φ,n) =R(−φ, −n)
  • R(π + φ,n) =R(π − φ, −n).

Using these properties one can show that any rotation can be represented by a unique angleφ in the range 0 ≤ φ ≤π and a unit vectorn such that

  • n is arbitrary ifφ = 0
  • n is unique if 0 <φ <π
  • n is unique up to asign ifφ =π (that is, the rotationsR(π, ±n) are identical).

In the next section, this representation of rotations is used to identify SO(3) topologically with three-dimensional real projective space.

Topology

edit

The Lie group SO(3) isdiffeomorphic to thereal projective spaceP3(R).{\displaystyle \mathbb {P} ^{3}(\mathbb {R} ).} [4]

Consider the solid ball inR3{\displaystyle \mathbb {R} ^{3}}  of radiusπ (that is, all points ofR3{\displaystyle \mathbb {R} ^{3}}  of distanceπ or less from the origin). Given the above, for every point in this ball there is a rotation, with axis through the point and the origin, and rotation angle equal to the distance of the point from the origin. The identity rotation corresponds to the point at the center of the ball. Rotations through an angle 𝜃 between 0 andπ (not including either) are on the same axis at the same distance. Rotation through angles between 0 and −π correspond to the point on the same axis and distance from the origin but on the opposite side of the origin. The one remaining issue is that the two rotations throughπ and through −π are the same. So weidentify (or "glue together")antipodal points on the surface of the ball. After this identification, we arrive at atopological spacehomeomorphic to the rotation group.

Indeed, the ball with antipodal surface points identified is asmooth manifold, and this manifold isdiffeomorphic to the rotation group. It is also diffeomorphic to thereal 3-dimensional projective spaceP3(R),{\displaystyle \mathbb {P} ^{3}(\mathbb {R} ),}  so the latter can also serve as a topological model for the rotation group.

These identifications illustrate that SO(3) isconnected but notsimply connected. As to the latter, in the ball with antipodal surface points identified, consider the path running from the "north pole" straight through the interior down to the south pole. This is a closed loop, since the north pole and the south pole are identified. This loop cannot be shrunk to a point, since no matter how it is deformed, the start and end point have to remain antipodal, or else the loop will "break open". In terms of rotations, this loop represents a continuous sequence of rotations about thez-axis starting (by example) at the identity (center of the ball), through the south pole, jumping to the north pole and ending again at the identity rotation (i.e., a series of rotation through an angleφ whereφ runs from 0 to2π).

Surprisingly, running through the path twice, i.e., running from the north pole down to the south pole, jumping back to the north pole (using the fact that north and south poles are identified), and then again running from the north pole down to the south pole, so thatφ runs from 0 to 4π, gives a closed loop whichcan be shrunk to a single point: first move the paths continuously to the ball's surface, still connecting north pole to south pole twice. The second path can then be mirrored over to the antipodal side without changing the path at all. Now we have an ordinary closed loop on the surface of the ball, connecting the north pole to itself along a great circle. This circle can be shrunk to the north pole without problems. Theplate trick and similar tricks demonstrate this practically.

The same argument can be performed in general, and it shows that thefundamental group of SO(3) is thecyclic group of order 2 (a fundamental group with two elements). Inphysics applications, the non-triviality (more than one element) of the fundamental group allows for the existence of objects known asspinors, and is an important tool in the development of thespin–statistics theorem.

Theuniversal cover of SO(3) is aLie group calledSpin(3). The group Spin(3) is isomorphic to thespecial unitary group SU(2); it is also diffeomorphic to the unit3-sphereS3 and can be understood as the group ofversors (quaternions withabsolute value 1). The connection between quaternions and rotations, commonly exploited incomputer graphics, is explained inquaternions and spatial rotations. The map fromS3 onto SO(3) that identifies antipodal points ofS3 is asurjectivehomomorphism of Lie groups, withkernel {±1}. Topologically, this map is a two-to-onecovering map. (See theplate trick.)

Connection between SO(3) and SU(2)

edit

In this section, we give two different constructions of a two-to-one andsurjectivehomomorphism of SU(2) onto SO(3).

Using quaternions of unit norm

edit

The groupSU(2) isisomorphic to thequaternions of unit norm via a map given by[5]q=a1+bi+cj+dk=α+βj[αββ¯α¯]=U{\displaystyle q=a\mathbf {1} +b\mathbf {i} +c\mathbf {j} +d\mathbf {k} =\alpha +\beta \mathbf {j} \leftrightarrow {\begin{bmatrix}\alpha &\beta \\-{\overline {\beta }}&{\overline {\alpha }}\end{bmatrix}}=U} restricted toa2+b2+c2+d2=|α|2+|β|2=1{\textstyle a^{2}+b^{2}+c^{2}+d^{2}=|\alpha |^{2}+|\beta |^{2}=1}  whereqH{\textstyle q\in \mathbb {H} } ,a,b,c,dR{\textstyle a,b,c,d\in \mathbb {R} } ,USU(2){\textstyle U\in \operatorname {SU} (2)} , andα=a+biC{\displaystyle \alpha =a+bi\in \mathbb {C} } ,β=c+diC{\displaystyle \beta =c+di\in \mathbb {C} } .

Let us now identifyR3{\displaystyle \mathbb {R} ^{3}}  with the span ofi,j,k{\displaystyle \mathbf {i} ,\mathbf {j} ,\mathbf {k} } . One can then verify that ifv{\displaystyle v}  is inR3{\displaystyle \mathbb {R} ^{3}}  andq{\displaystyle q}  is a unit quaternion, thenqvq1R3.{\displaystyle qvq^{-1}\in \mathbb {R} ^{3}.} 

Furthermore, the mapvqvq1{\displaystyle v\mapsto qvq^{-1}}  is a rotation ofR3.{\displaystyle \mathbb {R} ^{3}.}  Moreover,(q)v(q)1{\displaystyle (-q)v(-q)^{-1}}  is the same asqvq1{\displaystyle qvq^{-1}} . This means that there is a2:1 homomorphism from quaternions of unit norm to the 3D rotation groupSO(3).

One can work this homomorphism out explicitly: the unit quaternion,q, withq=w+xi+yj+zk,1=w2+x2+y2+z2,{\displaystyle {\begin{aligned}q&=w+x\mathbf {i} +y\mathbf {j} +z\mathbf {k} ,\\1&=w^{2}+x^{2}+y^{2}+z^{2},\end{aligned}}} is mapped to the rotation matrixQ=[12y22z22xy2zw2xz+2yw2xy+2zw12x22z22yz2xw2xz2yw2yz+2xw12x22y2].{\displaystyle Q={\begin{bmatrix}1-2y^{2}-2z^{2}&2xy-2zw&2xz+2yw\\2xy+2zw&1-2x^{2}-2z^{2}&2yz-2xw\\2xz-2yw&2yz+2xw&1-2x^{2}-2y^{2}\end{bmatrix}}.} 

This is a rotation around the vector(x,y,z) by an angle2θ, wherecosθ =w and|sinθ| = ‖(x,y,z). The proper sign forsinθ is implied, once the signs of the axis components are fixed. The2:1-nature is apparent since bothq andq map to the sameQ.

Using Möbius transformations

edit
 
Stereographic projection from the sphere of radius1/2 from the north pole(x,y,z) = (0, 0,1/2) onto the planeM given byz = −1/2 coordinatized by(ξ,η), here shown in cross section.

The general reference for this section isGelfand, Minlos & Shapiro (1963). The pointsP on the sphere

S={(x,y,z)R3:x2+y2+z2=14}{\displaystyle \mathbf {S} =\left\{(x,y,z)\in \mathbb {R} ^{3}:x^{2}+y^{2}+z^{2}={\frac {1}{4}}\right\}} 

can, barring the north poleN, be put into one-to-one bijection with pointsS(P) =P' on the planeM defined byz = −1/2, see figure. The mapS is calledstereographic projection.

Let the coordinates onM be(ξ,η). The lineL passing throughN andP can be parametrized as

L(t)=N+t(NP)=(0,0,12)+t((0,0,12)(x,y,z)),tR.{\displaystyle L(t)=N+t(N-P)=\left(0,0,{\frac {1}{2}}\right)+t\left(\left(0,0,{\frac {1}{2}}\right)-(x,y,z)\right),\quad t\in \mathbb {R} .} 

Demanding that thez-coordinate ofL(t0){\displaystyle L(t_{0})}  equals1/2, one finds

t0=1z12.{\displaystyle t_{0}={\frac {1}{z-{\frac {1}{2}}}}.} 

We haveL(t0)=(ξ,η,1/2).{\displaystyle L(t_{0})=(\xi ,\eta ,-1/2).}  Hence the map

{S:SMP=(x,y,z)P=(ξ,η)=(x12z,y12z)ζ=ξ+iη{\displaystyle {\begin{cases}S:\mathbf {S} \to M\\P=(x,y,z)\longmapsto P'=(\xi ,\eta )=\left({\frac {x}{{\frac {1}{2}}-z}},{\frac {y}{{\frac {1}{2}}-z}}\right)\equiv \zeta =\xi +i\eta \end{cases}}} 

where, for later convenience, the planeM is identified with the complex planeC.{\displaystyle \mathbb {C} .} 

For the inverse, writeL as

L=N+s(PN)=(0,0,12)+s((ξ,η,12)(0,0,12)),{\displaystyle L=N+s(P'-N)=\left(0,0,{\frac {1}{2}}\right)+s\left(\left(\xi ,\eta ,-{\frac {1}{2}}\right)-\left(0,0,{\frac {1}{2}}\right)\right),} 

and demandx2 +y2 +z2 =1/4 to finds =1/1 +ξ2 +η2 and thus

{S1:MSP=(ξ,η)P=(x,y,z)=(ξ1+ξ2+η2,η1+ξ2+η2,1+ξ2+η22+2ξ2+2η2){\displaystyle {\begin{cases}S^{-1}:M\to \mathbf {S} \\P'=(\xi ,\eta )\longmapsto P=(x,y,z)=\left({\frac {\xi }{1+\xi ^{2}+\eta ^{2}}},{\frac {\eta }{1+\xi ^{2}+\eta ^{2}}},{\frac {-1+\xi ^{2}+\eta ^{2}}{2+2\xi ^{2}+2\eta ^{2}}}\right)\end{cases}}} 

Ifg ∈ SO(3) is a rotation, then it will take points onS to points onS by its standard actionΠs(g) on the embedding spaceR3.{\displaystyle \mathbb {R} ^{3}.}  By composing this action withS one obtains a transformationS ∘ Πs(g) ∘S−1 ofM,

ζ=PPΠs(g)P=gPS(gP)Πu(g)ζ=ζ.{\displaystyle \zeta =P'\longmapsto P\longmapsto \Pi _{s}(g)P=gP\longmapsto S(gP)\equiv \Pi _{u}(g)\zeta =\zeta '.} 

ThusΠu(g) is a transformation ofC{\displaystyle \mathbb {C} }  associated to the transformationΠs(g) ofR3{\displaystyle \mathbb {R} ^{3}} .

It turns out thatg ∈ SO(3) represented in this way byΠu(g) can be expressed as a matrixΠu(g) ∈ SU(2) (where the notation is recycled to use the same name for the matrix as for the transformation ofC{\displaystyle \mathbb {C} }  it represents). To identify this matrix, consider first a rotationgφ about thez-axis through an angleφ,

x=xcosϕysinϕ,y=xsinϕ+ycosϕ,z=z.{\displaystyle {\begin{aligned}x'&=x\cos \phi -y\sin \phi ,\\y'&=x\sin \phi +y\cos \phi ,\\z'&=z.\end{aligned}}} 

Hence

ζ=x+iy12z=eiϕ(x+iy)12z=eiϕζ=eiϕ2ζ+00ζ+eiϕ2,{\displaystyle \zeta '={\frac {x'+iy'}{{\frac {1}{2}}-z'}}={\frac {e^{i\phi }(x+iy)}{{\frac {1}{2}}-z}}=e^{i\phi }\zeta ={\frac {e^{\frac {i\phi }{2}}\zeta +0}{0\zeta +e^{-{\frac {i\phi }{2}}}}},} 

which, unsurprisingly, is a rotation in the complex plane. In an analogous way, ifgθ is a rotation about thex-axis through an angleθ, then

w=eiθw,w=y+iz12x,{\displaystyle w'=e^{i\theta }w,\quad w={\frac {y+iz}{{\frac {1}{2}}-x}},} 

which, after a little algebra, becomes

ζ=cosθ2ζ+isinθ2isinθ2ζ+cosθ2.{\displaystyle \zeta '={\frac {\cos {\frac {\theta }{2}}\zeta +i\sin {\frac {\theta }{2}}}{i\sin {\frac {\theta }{2}}\zeta +\cos {\frac {\theta }{2}}}}.} 

These two rotations,gϕ,gθ,{\displaystyle g_{\phi },g_{\theta },}  thus correspond tobilinear transforms ofR2CM, namely, they are examples ofMöbius transformations.

A general Möbius transformation is given by

ζ=αζ+βγζ+δ,αδβγ0.{\displaystyle \zeta '={\frac {\alpha \zeta +\beta }{\gamma \zeta +\delta }},\quad \alpha \delta -\beta \gamma \neq 0.} 

The rotations,gϕ,gθ{\displaystyle g_{\phi },g_{\theta }}  generate all ofSO(3) and the composition rules of the Möbius transformations show that any composition ofgϕ,gθ{\displaystyle g_{\phi },g_{\theta }}  translates to the corresponding composition of Möbius transformations. The Möbius transformations can be represented by matrices

(αβγδ),αδβγ=1,{\displaystyle {\begin{pmatrix}\alpha &\beta \\\gamma &\delta \end{pmatrix}},\qquad \alpha \delta -\beta \gamma =1,} 

since a common factor ofα,β,γ,δ cancels.

For the same reason, the matrix isnot uniquely defined since multiplication byI has no effect on either the determinant or the Möbius transformation. The composition law of Möbius transformations follow that of the corresponding matrices. The conclusion is that each Möbius transformation corresponds to two matricesg, −g ∈ SL(2,C).

Using this correspondence one may write

Πu(gϕ)=Πu[(cosϕsinϕ0sinϕcosϕ0001)]=±(eiϕ200eiϕ2),Πu(gθ)=Πu[(1000cosθsinθ0sinθcosθ)]=±(cosθ2isinθ2isinθ2cosθ2).{\displaystyle {\begin{aligned}\Pi _{u}(g_{\phi })&=\Pi _{u}\left[{\begin{pmatrix}\cos \phi &-\sin \phi &0\\\sin \phi &\cos \phi &0\\0&0&1\end{pmatrix}}\right]=\pm {\begin{pmatrix}e^{i{\frac {\phi }{2}}}&0\\0&e^{-i{\frac {\phi }{2}}}\end{pmatrix}},\\\Pi _{u}(g_{\theta })&=\Pi _{u}\left[{\begin{pmatrix}1&0&0\\0&\cos \theta &-\sin \theta \\0&\sin \theta &\cos \theta \end{pmatrix}}\right]=\pm {\begin{pmatrix}\cos {\frac {\theta }{2}}&i\sin {\frac {\theta }{2}}\\i\sin {\frac {\theta }{2}}&\cos {\frac {\theta }{2}}\end{pmatrix}}.\end{aligned}}} 

These matrices are unitary and thusΠu(SO(3)) ⊂ SU(2) ⊂ SL(2,C). In terms ofEuler angles[nb 1] one finds for a general rotation

g(ϕ,θ,ψ)=gϕgθgψ=(cosϕsinϕ0sinϕcosϕ0001)(1000cosθsinθ0sinθcosθ)(cosψsinψ0sinψcosψ0001)=(cosϕcosψcosθsinϕsinψcosϕsinψcosθsinϕcosψsinϕsinθsinϕcosψ+cosθcosϕsinψsinϕsinψ+cosθcosϕcosψcosϕsinθsinψsinθcosψsinθcosθ),{\displaystyle {\begin{aligned}g(\phi ,\theta ,\psi )=g_{\phi }g_{\theta }g_{\psi }&={\begin{pmatrix}\cos \phi &-\sin \phi &0\\\sin \phi &\cos \phi &0\\0&0&1\end{pmatrix}}{\begin{pmatrix}1&0&0\\0&\cos \theta &-\sin \theta \\0&\sin \theta &\cos \theta \end{pmatrix}}{\begin{pmatrix}\cos \psi &-\sin \psi &0\\\sin \psi &\cos \psi &0\\0&0&1\end{pmatrix}}\\&={\begin{pmatrix}\cos \phi \cos \psi -\cos \theta \sin \phi \sin \psi &-\cos \phi \sin \psi -\cos \theta \sin \phi \cos \psi &\sin \phi \sin \theta \\\sin \phi \cos \psi +\cos \theta \cos \phi \sin \psi &-\sin \phi \sin \psi +\cos \theta \cos \phi \cos \psi &-\cos \phi \sin \theta \\\sin \psi \sin \theta &\cos \psi \sin \theta &\cos \theta \end{pmatrix}},\end{aligned}}} 1

one has[6]

Πu(g(ϕ,θ,ψ))=±(eiϕ200eiϕ2)(cosθ2isinθ2isinθ2cosθ2)(eiψ200eiψ2)=±(cosθ2eiϕ+ψ2isinθ2eiϕψ2isinθ2eiϕψ2cosθ2eiϕ+ψ2).{\displaystyle {\begin{aligned}\Pi _{u}(g(\phi ,\theta ,\psi ))&=\pm {\begin{pmatrix}e^{i{\frac {\phi }{2}}}&0\\0&e^{-i{\frac {\phi }{2}}}\end{pmatrix}}{\begin{pmatrix}\cos {\frac {\theta }{2}}&i\sin {\frac {\theta }{2}}\\i\sin {\frac {\theta }{2}}&\cos {\frac {\theta }{2}}\end{pmatrix}}{\begin{pmatrix}e^{i{\frac {\psi }{2}}}&0\\0&e^{-i{\frac {\psi }{2}}}\end{pmatrix}}\\&=\pm {\begin{pmatrix}\cos {\frac {\theta }{2}}e^{i{\frac {\phi +\psi }{2}}}&i\sin {\frac {\theta }{2}}e^{i{\frac {\phi -\psi }{2}}}\\i\sin {\frac {\theta }{2}}e^{-i{\frac {\phi -\psi }{2}}}&\cos {\frac {\theta }{2}}e^{-i{\frac {\phi +\psi }{2}}}\end{pmatrix}}.\end{aligned}}} 2

For the converse, consider a general matrix

±Πu(gα,β)=±(αββ¯α¯)SU(2).{\displaystyle \pm \Pi _{u}(g_{\alpha ,\beta })=\pm {\begin{pmatrix}\alpha &\beta \\-{\overline {\beta }}&{\overline {\alpha }}\end{pmatrix}}\in \operatorname {SU} (2).} 

Make the substitutions

cosθ2=|α|,sinθ2=|β|,(0θπ),ϕ+ψ2=argα,ψϕ2=argβ.{\displaystyle {\begin{aligned}\cos {\frac {\theta }{2}}&=|\alpha |,&\sin {\frac {\theta }{2}}&=|\beta |,&(0\leq \theta \leq \pi ),\\{\frac {\phi +\psi }{2}}&=\arg \alpha ,&{\frac {\psi -\phi }{2}}&=\arg \beta .&\end{aligned}}} 

With the substitutions,Π(gα,β) assumes the form of the right hand side (RHS) of (2), which corresponds underΠu to a matrix on the form of the RHS of (1) with the sameφ,θ,ψ. In terms of the complex parametersα,β,

gα,β=(12(α2β2+α2¯β2¯)i2(α2β2+α2¯+β2¯)αβα¯β¯i2(α2β2α2¯+β2¯)12(α2+β2+α2¯+β2¯)i(+αβα¯β¯)αβ¯+α¯βi(αβ¯+α¯β)αα¯ββ¯).{\displaystyle g_{\alpha ,\beta }={\begin{pmatrix}{\frac {1}{2}}\left(\alpha ^{2}-\beta ^{2}+{\overline {\alpha ^{2}}}-{\overline {\beta ^{2}}}\right)&{\frac {i}{2}}\left(-\alpha ^{2}-\beta ^{2}+{\overline {\alpha ^{2}}}+{\overline {\beta ^{2}}}\right)&-\alpha \beta -{\overline {\alpha }}{\overline {\beta }}\\{\frac {i}{2}}\left(\alpha ^{2}-\beta ^{2}-{\overline {\alpha ^{2}}}+{\overline {\beta ^{2}}}\right)&{\frac {1}{2}}\left(\alpha ^{2}+\beta ^{2}+{\overline {\alpha ^{2}}}+{\overline {\beta ^{2}}}\right)&-i\left(+\alpha \beta -{\overline {\alpha }}{\overline {\beta }}\right)\\\alpha {\overline {\beta }}+{\overline {\alpha }}\beta &i\left(-\alpha {\overline {\beta }}+{\overline {\alpha }}\beta \right)&\alpha {\overline {\alpha }}-\beta {\overline {\beta }}\end{pmatrix}}.} 

To verify this, substitute forα.β the elements of the matrix on the RHS of (2). After some manipulation, the matrix assumes the form of the RHS of (1).

It is clear from the explicit form in terms of Euler angles that the map

{p:SU(2)SO(3)±Πu(gαβ)gαβ{\displaystyle {\begin{cases}p:\operatorname {SU} (2)\to \operatorname {SO} (3)\\\pm \Pi _{u}(g_{\alpha \beta })\mapsto g_{\alpha \beta }\end{cases}}} 

just described is a smooth,2:1 and surjectivegroup homomorphism. It is hence an explicit description of theuniversal covering space ofSO(3) from theuniversal covering groupSU(2).

Lie algebra

edit

Associated with everyLie group is itsLie algebra, a linear space of the same dimension as the Lie group, closed under a bilinear alternating product called theLie bracket. The Lie algebra ofSO(3) is denoted byso(3){\displaystyle {\mathfrak {so}}(3)}  and consists of allskew-symmetric3 × 3 matrices.[7] This may be seen by differentiating theorthogonality condition,ATA =I,A ∈ SO(3).[nb 2] The Lie bracket of two elements ofso(3){\displaystyle {\mathfrak {so}}(3)}  is, as for the Lie algebra of every matrix group, given by the matrixcommutator,[A1,A2] =A1A2A2A1, which is again a skew-symmetric matrix. The Lie algebra bracket captures the essence of the Lie group product in a sense made precise by theBaker–Campbell–Hausdorff formula.

The elements ofso(3){\displaystyle {\mathfrak {so}}(3)}  are the "infinitesimal generators" of rotations, i.e., they are the elements of thetangent space of the manifold SO(3) at the identity element. IfR(ϕ,n){\displaystyle R(\phi ,{\boldsymbol {n}})}  denotes a counterclockwise rotation with angle φ about the axis specified by the unit vectorn,{\displaystyle {\boldsymbol {n}},}  then

uR3:ddϕ|ϕ=0R(ϕ,n)u=n×u.{\displaystyle \forall {\boldsymbol {u}}\in \mathbb {R} ^{3}:\qquad \left.{\frac {\operatorname {d} }{\operatorname {d} \phi }}\right|_{\phi =0}R(\phi ,{\boldsymbol {n}}){\boldsymbol {u}}={\boldsymbol {n}}\times {\boldsymbol {u}}.} 

This can be used to show that the Lie algebraso(3){\displaystyle {\mathfrak {so}}(3)}  (with commutator) is isomorphic to the Lie algebraR3{\displaystyle \mathbb {R} ^{3}}  (withcross product). Under this isomorphism, anEuler vectorωR3{\displaystyle {\boldsymbol {\omega }}\in \mathbb {R} ^{3}}  corresponds to the linear mapω~{\displaystyle {\widetilde {\boldsymbol {\omega }}}}  defined byω~(u)=ω×u.{\displaystyle {\widetilde {\boldsymbol {\omega }}}({\boldsymbol {u}})={\boldsymbol {\omega }}\times {\boldsymbol {u}}.} 

In more detail, most often a suitable basis forso(3){\displaystyle {\mathfrak {so}}(3)}  as a3-dimensional vector space is

Lx=[000001010],Ly=[001000100],Lz=[010100000].{\displaystyle {\boldsymbol {L}}_{x}={\begin{bmatrix}0&0&0\\0&0&-1\\0&1&0\end{bmatrix}},\quad {\boldsymbol {L}}_{y}={\begin{bmatrix}0&0&1\\0&0&0\\-1&0&0\end{bmatrix}},\quad {\boldsymbol {L}}_{z}={\begin{bmatrix}0&-1&0\\1&0&0\\0&0&0\end{bmatrix}}.} 

Thecommutation relations of these basis elements are,

[Lx,Ly]=Lz,[Lz,Lx]=Ly,[Ly,Lz]=Lx{\displaystyle [{\boldsymbol {L}}_{x},{\boldsymbol {L}}_{y}]={\boldsymbol {L}}_{z},\quad [{\boldsymbol {L}}_{z},{\boldsymbol {L}}_{x}]={\boldsymbol {L}}_{y},\quad [{\boldsymbol {L}}_{y},{\boldsymbol {L}}_{z}]={\boldsymbol {L}}_{x}} 

which agree with the relations of the threestandard unit vectors ofR3{\displaystyle \mathbb {R} ^{3}}  under the cross product.

As announced above, one can identify any matrix in this Lie algebra with an Euler vectorω=(x,y,z)R3,{\displaystyle {\boldsymbol {\omega }}=(x,y,z)\in \mathbb {R} ^{3},} [8]

ω^=ωL=xLx+yLy+zLz=[0zyz0xyx0]so(3).{\displaystyle {\widehat {\boldsymbol {\omega }}}={\boldsymbol {\omega }}\cdot {\boldsymbol {L}}=x{\boldsymbol {L}}_{x}+y{\boldsymbol {L}}_{y}+z{\boldsymbol {L}}_{z}={\begin{bmatrix}0&-z&y\\z&0&-x\\-y&x&0\end{bmatrix}}\in {\mathfrak {so}}(3).} 

This identification is sometimes called thehat-map.[9] Under this identification, theso(3){\displaystyle {\mathfrak {so}}(3)}  bracket corresponds inR3{\displaystyle \mathbb {R} ^{3}}  to thecross product,

[u^,v^]=u×v^.{\displaystyle \left[{\widehat {\boldsymbol {u}}},{\widehat {\boldsymbol {v}}}\right]={\widehat {{\boldsymbol {u}}\times {\boldsymbol {v}}}}.} 

The matrix identified with a vectoru{\displaystyle {\boldsymbol {u}}}  has the property that

u^v=u×v,{\displaystyle {\widehat {\boldsymbol {u}}}{\boldsymbol {v}}={\boldsymbol {u}}\times {\boldsymbol {v}},} 

where the left-hand side we have ordinary matrix multiplication. This impliesu{\displaystyle {\boldsymbol {u}}}  is in thenull space of the skew-symmetric matrix with which it is identified, becauseu×u=0.{\displaystyle {\boldsymbol {u}}\times {\boldsymbol {u}}={\boldsymbol {0}}.} 

A note on Lie algebras

edit

InLie algebra representations, the group SO(3) is compact and simple of rank 1, and so it has a single independentCasimir element, a quadratic invariant function of the three generators which commutes with all of them. The Killing form for the rotation group is just theKronecker delta, and so this Casimir invariant is simply the sum of the squares of the generators,Jx,Jy,Jz,{\displaystyle {\boldsymbol {J}}_{x},{\boldsymbol {J}}_{y},{\boldsymbol {J}}_{z},}  of the algebra

[Jx,Jy]=Jz,[Jz,Jx]=Jy,[Jy,Jz]=Jx.{\displaystyle [{\boldsymbol {J}}_{x},{\boldsymbol {J}}_{y}]={\boldsymbol {J}}_{z},\quad [{\boldsymbol {J}}_{z},{\boldsymbol {J}}_{x}]={\boldsymbol {J}}_{y},\quad [{\boldsymbol {J}}_{y},{\boldsymbol {J}}_{z}]={\boldsymbol {J}}_{x}.} 

That is, the Casimir invariant is given by

J2JJ=Jx2+Jy2+Jz2I.{\displaystyle {\boldsymbol {J}}^{2}\equiv {\boldsymbol {J}}\cdot {\boldsymbol {J}}={\boldsymbol {J}}_{x}^{2}+{\boldsymbol {J}}_{y}^{2}+{\boldsymbol {J}}_{z}^{2}\propto {\boldsymbol {I}}.} 

For unitary irreduciblerepresentationsDj, the eigenvalues of this invariant are real and discrete, and characterize each representation, which is finite dimensional, of dimensionality2j+1{\displaystyle 2j+1} . That is, the eigenvalues of this Casimir operator are

J2=j(j+1)I2j+1,{\displaystyle {\boldsymbol {J}}^{2}=-j(j+1){\boldsymbol {I}}_{2j+1},} 

wherej is integer or half-integer, and referred to as thespin orangular momentum.

So, the 3 × 3 generatorsL displayed above act on the triplet (spin 1) representation, while the 2 × 2 generators below,t, act on thedoublet (spin-1/2) representation. By takingKronecker products ofD1/2 with itself repeatedly, one may construct all higher irreducible representationsDj. That is, the resulting generators for higher spin systems in three spatial dimensions, for arbitrarily largej, can be calculated using thesespin operators andladder operators.

For every unitary irreducible representationsDj there is an equivalent one,Dj−1. All infinite-dimensional irreducible representations must be non-unitary, since the group is compact.

Inquantum mechanics, the Casimir invariant is the "angular-momentum-squared" operator; integer values of spinj characterizebosonic representations, while half-integer valuesfermionic representations. Theantihermitian matrices used above are utilized asspin operators, after they are multiplied byi, so they are nowhermitian (like the Pauli matrices). Thus, in this language,

[Jx,Jy]=iJz,[Jz,Jx]=iJy,[Jy,Jz]=iJx.{\displaystyle [{\boldsymbol {J}}_{x},{\boldsymbol {J}}_{y}]=i{\boldsymbol {J}}_{z},\quad [{\boldsymbol {J}}_{z},{\boldsymbol {J}}_{x}]=i{\boldsymbol {J}}_{y},\quad [{\boldsymbol {J}}_{y},{\boldsymbol {J}}_{z}]=i{\boldsymbol {J}}_{x}.} 

and hence

J2=j(j+1)I2j+1.{\displaystyle {\boldsymbol {J}}^{2}=j(j+1){\boldsymbol {I}}_{2j+1}.} 

Explicit expressions for theseDj are,

(Jz(j))ba=(j+1a)δb,a(Jx(j))ba=12(δb,a+1+δb+1,a)(j+1)(a+b1)ab(Jy(j))ba=12i(δb,a+1δb+1,a)(j+1)(a+b1)ab{\displaystyle {\begin{aligned}\left({\boldsymbol {J}}_{z}^{(j)}\right)_{ba}&=(j+1-a)\delta _{b,a}\\\left({\boldsymbol {J}}_{x}^{(j)}\right)_{ba}&={\frac {1}{2}}\left(\delta _{b,a+1}+\delta _{b+1,a}\right){\sqrt {(j+1)(a+b-1)-ab}}\\\left({\boldsymbol {J}}_{y}^{(j)}\right)_{ba}&={\frac {1}{2i}}\left(\delta _{b,a+1}-\delta _{b+1,a}\right){\sqrt {(j+1)(a+b-1)-ab}}\\\end{aligned}}} 

wherej is arbitrary and1a,b2j+1{\displaystyle 1\leq a,b\leq 2j+1} .

For example, the resulting spin matrices for spin 1 (j=1{\displaystyle j=1} ) are

Jx=12(010101010)Jy=12(0i0i0i0i0)Jz=(100000001){\displaystyle {\begin{aligned}{\boldsymbol {J}}_{x}&={\frac {1}{\sqrt {2}}}{\begin{pmatrix}0&1&0\\1&0&1\\0&1&0\end{pmatrix}}\\{\boldsymbol {J}}_{y}&={\frac {1}{\sqrt {2}}}{\begin{pmatrix}0&-i&0\\i&0&-i\\0&i&0\end{pmatrix}}\\{\boldsymbol {J}}_{z}&={\begin{pmatrix}1&0&0\\0&0&0\\0&0&-1\end{pmatrix}}\end{aligned}}} 

Note, however, how these are in an equivalent, but different basis, thespherical basis, than the aboveiL in the Cartesian basis.[nb 3]

For higher spins, such as spin3/2 (j=32{\displaystyle j={\tfrac {3}{2}}} ):

Jx=12(0300302002030030)Jy=12(0i300i302i002i0i300i30)Jz=12(3000010000100003).{\displaystyle {\begin{aligned}{\boldsymbol {J}}_{x}&={\frac {1}{2}}{\begin{pmatrix}0&{\sqrt {3}}&0&0\\{\sqrt {3}}&0&2&0\\0&2&0&{\sqrt {3}}\\0&0&{\sqrt {3}}&0\end{pmatrix}}\\{\boldsymbol {J}}_{y}&={\frac {1}{2}}{\begin{pmatrix}0&-i{\sqrt {3}}&0&0\\i{\sqrt {3}}&0&-2i&0\\0&2i&0&-i{\sqrt {3}}\\0&0&i{\sqrt {3}}&0\end{pmatrix}}\\{\boldsymbol {J}}_{z}&={\frac {1}{2}}{\begin{pmatrix}3&0&0&0\\0&1&0&0\\0&0&-1&0\\0&0&0&-3\end{pmatrix}}.\end{aligned}}} 

For spin5/2 (j=52{\displaystyle j={\tfrac {5}{2}}} ),

Jx=12(0500005022000022030000302200002205000050)Jy=12(0i50000i502i200002i203i00003i02i200002i20i50000i50)Jz=12(500000030000001000000100000030000005).{\displaystyle {\begin{aligned}{\boldsymbol {J}}_{x}&={\frac {1}{2}}{\begin{pmatrix}0&{\sqrt {5}}&0&0&0&0\\{\sqrt {5}}&0&2{\sqrt {2}}&0&0&0\\0&2{\sqrt {2}}&0&3&0&0\\0&0&3&0&2{\sqrt {2}}&0\\0&0&0&2{\sqrt {2}}&0&{\sqrt {5}}\\0&0&0&0&{\sqrt {5}}&0\end{pmatrix}}\\{\boldsymbol {J}}_{y}&={\frac {1}{2}}{\begin{pmatrix}0&-i{\sqrt {5}}&0&0&0&0\\i{\sqrt {5}}&0&-2i{\sqrt {2}}&0&0&0\\0&2i{\sqrt {2}}&0&-3i&0&0\\0&0&3i&0&-2i{\sqrt {2}}&0\\0&0&0&2i{\sqrt {2}}&0&-i{\sqrt {5}}\\0&0&0&0&i{\sqrt {5}}&0\end{pmatrix}}\\{\boldsymbol {J}}_{z}&={\frac {1}{2}}{\begin{pmatrix}5&0&0&0&0&0\\0&3&0&0&0&0\\0&0&1&0&0&0\\0&0&0&-1&0&0\\0&0&0&0&-3&0\\0&0&0&0&0&-5\end{pmatrix}}.\end{aligned}}} 

Isomorphism with 𝖘𝖚(2)

edit

The Lie algebrasso(3){\displaystyle {\mathfrak {so}}(3)}  andsu(2){\displaystyle {\mathfrak {su}}(2)}  are isomorphic. One basis forsu(2){\displaystyle {\mathfrak {su}}(2)}  is given by[10]

t1=12[0ii0],t2=12[0110],t3=12[i00i].{\displaystyle {\boldsymbol {t}}_{1}={\frac {1}{2}}{\begin{bmatrix}0&-i\\-i&0\end{bmatrix}},\quad {\boldsymbol {t}}_{2}={\frac {1}{2}}{\begin{bmatrix}0&-1\\1&0\end{bmatrix}},\quad {\boldsymbol {t}}_{3}={\frac {1}{2}}{\begin{bmatrix}-i&0\\0&i\end{bmatrix}}.} 

These are related to thePauli matrices by

ti12iσi.{\displaystyle {\boldsymbol {t}}_{i}\longleftrightarrow {\frac {1}{2i}}\sigma _{i}.} 

The Pauli matrices abide by the physicists' convention for Lie algebras. In that convention, Lie algebra elements are multiplied byi, the exponential map (below) is defined with an extra factor ofi in the exponent and thestructure constants remain the same, but thedefinition of them acquires a factor ofi. Likewise, commutation relations acquire a factor ofi. The commutation relations for theti{\displaystyle {\boldsymbol {t}}_{i}}  are

[ti,tj]=εijktk,{\displaystyle [{\boldsymbol {t}}_{i},{\boldsymbol {t}}_{j}]=\varepsilon _{ijk}{\boldsymbol {t}}_{k},} 

whereεijk is the totally anti-symmetric symbol withε123 = 1. The isomorphism betweenso(3){\displaystyle {\mathfrak {so}}(3)}  andsu(2){\displaystyle {\mathfrak {su}}(2)}  can be set up in several ways. For later convenience,so(3){\displaystyle {\mathfrak {so}}(3)}  andsu(2){\displaystyle {\mathfrak {su}}(2)}  are identified by mapping

Lxt1,Lyt2,Lzt3,{\displaystyle {\boldsymbol {L}}_{x}\longleftrightarrow {\boldsymbol {t}}_{1},\quad {\boldsymbol {L}}_{y}\longleftrightarrow {\boldsymbol {t}}_{2},\quad {\boldsymbol {L}}_{z}\longleftrightarrow {\boldsymbol {t}}_{3},} 

and extending by linearity.

Exponential map

edit

The exponential map forSO(3), is, sinceSO(3) is a matrix Lie group, defined using the standardmatrix exponential series,

{exp:so(3)SO(3)AeA=k=01k!Ak=I+A+12A2+.{\displaystyle {\begin{cases}\exp :{\mathfrak {so}}(3)\to \operatorname {SO} (3)\\A\mapsto e^{A}=\sum _{k=0}^{\infty }{\frac {1}{k!}}A^{k}=I+A+{\tfrac {1}{2}}A^{2}+\cdots .\end{cases}}} 

For anyskew-symmetric matrixA ∈ 𝖘𝖔(3),eA is always inSO(3). The proof uses the elementary properties of the matrix exponential

(eA)TeA=eATeA=eAT+A=eA+A=eAA=eA(eA)T=e0=I.{\displaystyle \left(e^{A}\right)^{\textsf {T}}e^{A}=e^{A^{\textsf {T}}}e^{A}=e^{A^{\textsf {T}}+A}=e^{-A+A}=e^{A-A}=e^{A}\left(e^{A}\right)^{\textsf {T}}=e^{0}=I.} 

since the matricesA andAT commute, this can be easily proven with the skew-symmetric matrix condition. This is not enough to show that𝖘𝖔(3) is the corresponding Lie algebra forSO(3), and shall be proven separately.

The level of difficulty of proof depends on how a matrix group Lie algebra is defined.Hall (2003) defines the Lie algebra as the set of matrices

{AM(n,R)|etASO(3)t},{\displaystyle \left\{A\in \operatorname {M} (n,\mathbb {R} )\left|e^{tA}\in \operatorname {SO} (3)\forall t\right.\right\},} 

in which case it is trivial.Rossmann (2002) uses for a definition derivatives of smooth curve segments inSO(3) through the identity taken at the identity, in which case it is harder.[11]

For a fixedA ≠ 0,etA, −∞ <t < ∞ is aone-parameter subgroup along ageodesic inSO(3). That this gives a one-parameter subgroup follows directly from properties of the exponential map.[12]

The exponential map provides adiffeomorphism between a neighborhood of the origin in the𝖘𝖔(3) and a neighborhood of the identity in theSO(3).[13] For a proof, seeClosed subgroup theorem.

The exponential map issurjective. This follows from the fact that everyR ∈ SO(3), since every rotation leaves an axis fixed (Euler's rotation theorem), and is conjugate to ablock diagonal matrix of the form

D=(cosθsinθ0sinθcosθ0001)=eθLz,{\displaystyle D={\begin{pmatrix}\cos \theta &-\sin \theta &0\\\sin \theta &\cos \theta &0\\0&0&1\end{pmatrix}}=e^{\theta L_{z}},} 

such thatA =BDB−1, and that

BeθLzB1=eBθLzB1,{\displaystyle Be^{\theta L_{z}}B^{-1}=e^{B\theta L_{z}B^{-1}},} 

together with the fact that𝖘𝖔(3) is closed under theadjoint action ofSO(3), meaning thatBθLzB−1 ∈ 𝖘𝖔(3).

Thus, e.g., it is easy to check the popular identity

eπLx/2eθLzeπLx/2=eθLy.{\displaystyle e^{-\pi L_{x}/2}e^{\theta L_{z}}e^{\pi L_{x}/2}=e^{\theta L_{y}}.} 

As shown above, every elementA ∈ 𝖘𝖔(3) is associated with a vectorω =θu, whereu = (x,y,z) is a unit magnitude vector. Sinceu is in the null space ofA, if one now rotates to a new basis, through some other orthogonal matrixO, withu as thez axis, the final column and row of the rotation matrix in the new basis will be zero.

Thus, we know in advance from the formula for the exponential thatexp(OAOT) must leaveu fixed. It is mathematically impossible to supply a straightforward formula for such a basis as a function ofu, because its existence would violate thehairy ball theorem; but direct exponentiation is possible, andyields

exp(ω~)=exp(θ(uL))=exp(θ[0zyz0xyx0])=I+2cs(uL)+2s2(uL)2=[2(x21)s2+12xys22zcs2xzs2+2ycs2xys2+2zcs2(y21)s2+12yzs22xcs2xzs22ycs2yzs2+2xcs2(z21)s2+1],{\displaystyle {\begin{aligned}\exp({\tilde {\boldsymbol {\omega }}})&=\exp(\theta ({\boldsymbol {u\cdot L}}))=\exp \left(\theta {\begin{bmatrix}0&-z&y\\z&0&-x\\-y&x&0\end{bmatrix}}\right)\\[4pt]&={\boldsymbol {I}}+2cs({\boldsymbol {u\cdot L}})+2s^{2}({\boldsymbol {u\cdot L}})^{2}\\[4pt]&={\begin{bmatrix}2\left(x^{2}-1\right)s^{2}+1&2xys^{2}-2zcs&2xzs^{2}+2ycs\\2xys^{2}+2zcs&2\left(y^{2}-1\right)s^{2}+1&2yzs^{2}-2xcs\\2xzs^{2}-2ycs&2yzs^{2}+2xcs&2\left(z^{2}-1\right)s^{2}+1\end{bmatrix}},\end{aligned}}} 

wherec=cosθ2{\textstyle c=\cos {\frac {\theta }{2}}}  ands=sinθ2{\textstyle s=\sin {\frac {\theta }{2}}} . This is recognized as a matrix for a rotation around axisu by the angleθ: cf.Rodrigues' rotation formula.

Logarithm map

edit

GivenR ∈ SO(3), letA=12(RRT){\displaystyle A={\tfrac {1}{2}}\left(R-R^{\mathrm {T} }\right)}  denote the antisymmetric part and letA=12Tr(A2).{\textstyle \|A\|={\sqrt {-{\frac {1}{2}}\operatorname {Tr} \left(A^{2}\right)}}.}  Then, the logarithm ofR is given by[9]

logR=sin1AAA.{\displaystyle \log R={\frac {\sin ^{-1}\|A\|}{\|A\|}}A.} 

This is manifest by inspection of the mixed symmetry form of Rodrigues' formula,

eX=I+sinθθX+2sin2θ2θ2X2,θ=X,{\displaystyle e^{X}=I+{\frac {\sin \theta }{\theta }}X+2{\frac {\sin ^{2}{\frac {\theta }{2}}}{\theta ^{2}}}X^{2},\quad \theta =\|X\|,} 

where the first and last term on the right-hand side are symmetric.

Uniform random sampling

edit

SO(3){\displaystyle SO(3)}  is doubly covered by the group of unit quaternions, which is isomorphic to the 3-sphere. Since theHaar measure on the unit quaternions is just the 3-area measure in 4 dimensions, the Haar measure onSO(3){\displaystyle SO(3)}  is just the pushforward of the 3-area measure.

Consequently, generating a uniformly random rotation inR3{\displaystyle \mathbb {R} ^{3}}  is equivalent to generating a uniformly random point on the 3-sphere. This can be accomplished by the following(1u1sin(2πu2),1u1cos(2πu2),u1sin(2πu3),u1cos(2πu3)){\displaystyle ({\sqrt {1-u_{1}}}\sin(2\pi u_{2}),{\sqrt {1-u_{1}}}\cos(2\pi u_{2}),{\sqrt {u_{1}}}\sin(2\pi u_{3}),{\sqrt {u_{1}}}\cos(2\pi u_{3}))} 

whereu1,u2,u3{\displaystyle u_{1},u_{2},u_{3}}  are uniformly random samples of[0,1]{\displaystyle [0,1]} .[14]

Baker–Campbell–Hausdorff formula

edit

SupposeX andY in the Lie algebra are given. Their exponentials,exp(X) andexp(Y), are rotation matrices, which can be multiplied. Since the exponential map is a surjection, for someZ in the Lie algebra,exp(Z) = exp(X) exp(Y), and one may tentatively write

Z=C(X,Y),{\displaystyle Z=C(X,Y),} 

forC some expression inX andY. Whenexp(X) andexp(Y) commute, thenZ =X +Y, mimicking the behavior of complex exponentiation.

The general case is given by the more elaborateBCH formula, a series expansion of nested Lie brackets.[15] For matrices, the Lie bracket is the same operation as thecommutator, which monitors lack of commutativity in multiplication. This general expansion unfolds as follows,[nb 4]

Z=C(X,Y)=X+Y+12[X,Y]+112[X,[X,Y]]112[Y,[X,Y]]+.{\displaystyle Z=C(X,Y)=X+Y+{\frac {1}{2}}[X,Y]+{\tfrac {1}{12}}[X,[X,Y]]-{\frac {1}{12}}[Y,[X,Y]]+\cdots .} 

The infinite expansion in the BCH formula forSO(3) reduces to a compact form,

Z=αX+βY+γ[X,Y],{\displaystyle Z=\alpha X+\beta Y+\gamma [X,Y],} 

for suitable trigonometric function coefficients(α,β,γ).

The trigonometric coefficients

The(α,β,γ) are given by

α=ϕcot(ϕ2)γ,β=θcot(θ2)γ,γ=sin1ddcθϕ,{\displaystyle \alpha =\phi \cot \left({\frac {\phi }{2}}\right)\gamma ,\qquad \beta =\theta \cot \left({\frac {\theta }{2}}\right)\gamma ,\qquad \gamma ={\frac {\sin ^{-1}d}{d}}{\frac {c}{\theta \phi }},} 

where

c=12sinθsinϕ2sin2θ2sin2ϕ2cos((u,v)),a=ccot(ϕ2),b=ccot(θ2),d=a2+b2+2abcos((u,v))+c2sin2((u,v)),{\displaystyle {\begin{aligned}c&={\frac {1}{2}}\sin \theta \sin \phi -2\sin ^{2}{\frac {\theta }{2}}\sin ^{2}{\frac {\phi }{2}}\cos(\angle (u,v)),\quad a=c\cot \left({\frac {\phi }{2}}\right),\quad b=c\cot \left({\frac {\theta }{2}}\right),\\d&={\sqrt {a^{2}+b^{2}+2ab\cos(\angle (u,v))+c^{2}\sin ^{2}(\angle (u,v))}},\end{aligned}}} 

for

θ=X,ϕ=Y,(u,v)=cos1X,YXY.{\displaystyle \theta =\|X\|,\quad \phi =\|Y\|,\quad \angle (u,v)=\cos ^{-1}{\frac {\langle X,Y\rangle }{\|X\|\|Y\|}}.} 

The inner product is theHilbert–Schmidt inner product and the norm is the associated norm. Under the hat-isomorphism,

u,v=12TrXTY,{\displaystyle \langle u,v\rangle ={\frac {1}{2}}\operatorname {Tr} X^{\mathrm {T} }Y,} 
which explains the factors forθ andφ. This drops out in the expression for the angle.

It is worthwhile to write this composite rotation generator as

αX+βY+γ[X,Y]=so(3)X+Y+12[X,Y]+112[X,[X,Y]]112[Y,[X,Y]]+,{\displaystyle \alpha X+\beta Y+\gamma [X,Y]{\underset {{\mathfrak {so}}(3)}{=}}X+Y+{\frac {1}{2}}[X,Y]+{\frac {1}{12}}[X,[X,Y]]-{\frac {1}{12}}[Y,[X,Y]]+\cdots ,} 

to emphasize that this is aLie algebra identity.

The above identity holds for allfaithful representations of𝖘𝖔(3). Thekernel of a Lie algebra homomorphism is anideal, but𝖘𝖔(3), beingsimple, has no nontrivial ideals and all nontrivial representations are hence faithful. It holds in particular in the doublet or spinor representation. The same explicit formula thus follows in a simpler way through Pauli matrices, cf. the2×2 derivation for SU(2).

The SU(2) case

ThePauli vector version of the same BCH formula is the somewhat simpler group composition law of SU(2),

eia(u^σ)eib(v^σ)=exp(csincsinasinb((icotbu^+icotav^)σ+12[iu^σ,iv^σ])),{\displaystyle e^{ia'\left({\hat {u}}\cdot {\vec {\sigma }}\right)}e^{ib'\left({\hat {v}}\cdot {\vec {\sigma }}\right)}=\exp \left({\frac {c'}{\sin c'}}\sin a'\sin b'\left(\left(i\cot b'{\hat {u}}+i\cot a'{\hat {v}}\right)\cdot {\vec {\sigma }}+{\frac {1}{2}}\left[i{\hat {u}}\cdot {\vec {\sigma }},i{\hat {v}}\cdot {\vec {\sigma }}\right]\right)\right),} 

where

cosc=cosacosbu^v^sinasinb,{\displaystyle \cos c'=\cos a'\cos b'-{\hat {u}}\cdot {\hat {v}}\sin a'\sin b',} 

thespherical law of cosines. (Note a', b', c' are angles, not thea,b,c above.)

This is manifestly of the same format as above,

Z=αX+βY+γ[X,Y],{\displaystyle Z=\alpha 'X+\beta 'Y+\gamma '[X,Y],} 

with

X=iau^σ,Y=ibv^σsu(2),{\displaystyle X=ia'{\hat {u}}\cdot \mathbf {\sigma } ,\quad Y=ib'{\hat {v}}\cdot \mathbf {\sigma } \in {\mathfrak {su}}(2),} 

so that

α=csincsinaacosbβ=csincsinbbcosaγ=12csincsinaasinbb.{\displaystyle {\begin{aligned}\alpha '&={\frac {c'}{\sin c'}}{\frac {\sin a'}{a'}}\cos b'\\\beta '&={\frac {c'}{\sin c'}}{\frac {\sin b'}{b'}}\cos a'\\\gamma '&={\frac {1}{2}}{\frac {c'}{\sin c'}}{\frac {\sin a'}{a'}}{\frac {\sin b'}{b'}}.\end{aligned}}} 

For uniform normalization of the generators in the Lie algebra involved, express the Pauli matrices in terms oft-matrices,σ → 2it, so that

aθ2,bϕ2.{\displaystyle a'\mapsto -{\frac {\theta }{2}},\quad b'\mapsto -{\frac {\phi }{2}}.} 

To verify then these are the same coefficients as above, compute the ratios of the coefficients,

αγ=θcotθ2=αγβγ=ϕcotϕ2=βγ.{\displaystyle {\begin{aligned}{\frac {\alpha '}{\gamma '}}&=\theta \cot {\frac {\theta }{2}}&={\frac {\alpha }{\gamma }}\\{\frac {\beta '}{\gamma '}}&=\phi \cot {\frac {\phi }{2}}&={\frac {\beta }{\gamma }}.\end{aligned}}} 

Finally,γ =γ' given the identityd = sin 2c'.

For the generaln ×n case, one might use Ref.[16]

The quaternion case

Thequaternion formulation of the composition of two rotations RB and RA also yields directly therotation axis and angle of the composite rotation RC = RBRA.

Let the quaternion associated with a spatial rotation R is constructed from itsrotation axisS and the rotation angleφ this axis. The associated quaternion is given by,

S=cosϕ2+sinϕ2S.{\displaystyle S=\cos {\frac {\phi }{2}}+\sin {\frac {\phi }{2}}\mathbf {S} .} 

Then the composition of the rotation RR with RA is the rotation RC = RBRA with rotation axis and angle defined by the product of the quaternions

A=cosα2+sinα2A and B=cosβ2+sinβ2B,{\displaystyle A=\cos {\frac {\alpha }{2}}+\sin {\frac {\alpha }{2}}\mathbf {A} \quad {\text{ and }}\quad B=\cos {\frac {\beta }{2}}+\sin {\frac {\beta }{2}}\mathbf {B} ,} 

that is

C=cosγ2+sinγ2C=(cosβ2+sinβ2B)(cosα2+sinα2A).{\displaystyle C=\cos {\frac {\gamma }{2}}+\sin {\frac {\gamma }{2}}\mathbf {C} =\left(\cos {\frac {\beta }{2}}+\sin {\frac {\beta }{2}}\mathbf {B} \right)\left(\cos {\frac {\alpha }{2}}+\sin {\frac {\alpha }{2}}\mathbf {A} \right).} 

Expand this product to obtain

cosγ2+sinγ2C=(cosβ2cosα2sinβ2sinα2BA)+(sinβ2cosα2B+sinα2cosβ2A+sinβ2sinα2B×A).{\displaystyle \cos {\frac {\gamma }{2}}+\sin {\frac {\gamma }{2}}\mathbf {C} =\left(\cos {\frac {\beta }{2}}\cos {\frac {\alpha }{2}}-\sin {\frac {\beta }{2}}\sin {\frac {\alpha }{2}}\mathbf {B} \cdot \mathbf {A} \right)+\left(\sin {\frac {\beta }{2}}\cos {\frac {\alpha }{2}}\mathbf {B} +\sin {\frac {\alpha }{2}}\cos {\frac {\beta }{2}}\mathbf {A} +\sin {\frac {\beta }{2}}\sin {\frac {\alpha }{2}}\mathbf {B} \times \mathbf {A} \right).} 

Divide both sides of this equation by the identity, which is thelaw of cosines on a sphere,

cosγ2=cosβ2cosα2sinβ2sinα2BA,{\displaystyle \cos {\frac {\gamma }{2}}=\cos {\frac {\beta }{2}}\cos {\frac {\alpha }{2}}-\sin {\frac {\beta }{2}}\sin {\frac {\alpha }{2}}\mathbf {B} \cdot \mathbf {A} ,} 

and compute

tanγ2C=tanβ2B+tanα2A+tanβ2tanα2B×A1tanβ2tanα2BA.{\displaystyle \tan {\frac {\gamma }{2}}\mathbf {C} ={\frac {\tan {\frac {\beta }{2}}\mathbf {B} +\tan {\frac {\alpha }{2}}\mathbf {A} +\tan {\frac {\beta }{2}}\tan {\frac {\alpha }{2}}\mathbf {B} \times \mathbf {A} }{1-\tan {\frac {\beta }{2}}\tan {\frac {\alpha }{2}}\mathbf {B} \cdot \mathbf {A} }}.} 

This is Rodrigues' formula for the axis of a composite rotation defined in terms of the axes of the two rotations. He derived this formula in 1840 (see page 408).[17]

The three rotation axesA,B, andC form a spherical triangle and the dihedral angles between the planes formed by the sides of this triangle are defined by the rotation angles.

Infinitesimal rotations

edit
This section is an excerpt fromInfinitesimal rotation matrix.[edit]

Aninfinitesimal rotation matrix or differential rotation matrix is amatrix representing aninfinitely smallrotation.

While arotation matrix is anorthogonal matrixRT=R1{\displaystyle R^{\mathsf {T}}=R^{-1}}  representing an element ofSO(n){\displaystyle SO(n)}  (thespecial orthogonal group), thedifferential of a rotation is askew-symmetric matrixAT=A{\displaystyle A^{\mathsf {T}}=-A}  in thetangent spaceso(n){\displaystyle {\mathfrak {so}}(n)}  (thespecial orthogonal Lie algebra), which is not itself a rotation matrix.

An infinitesimal rotation matrix has the form

I+dθA,{\displaystyle I+d\theta \,A,} 

whereI{\displaystyle I}  is the identity matrix,dθ{\displaystyle d\theta }  is vanishingly small, andAso(n).{\displaystyle A\in {\mathfrak {so}}(n).} 

For example, ifA=Lx,{\displaystyle A=L_{x},}  representing an infinitesimal three-dimensional rotation about thex-axis, a basis element ofso(3),{\displaystyle {\mathfrak {so}}(3),}  then

Lx=[000001010]{\displaystyle L_{x}={\begin{bmatrix}0&0&0\\0&0&-1\\0&1&0\end{bmatrix}}} ,

and

I+dθLx=[10001dθ0dθ1].{\displaystyle I+d\theta L_{x}={\begin{bmatrix}1&0&0\\0&1&-d\theta \\0&d\theta &1\end{bmatrix}}.} 
The computation rules for infinitesimal rotation matrices are the usual ones except that infinitesimals of second order are dropped. With these rules, these matrices do not satisfy all the same properties as ordinary finite rotation matrices under the usual treatment of infinitesimals.[18] It turns out thatthe order in which infinitesimal rotations are applied is irrelevant.

Realizations of rotations

edit

We have seen that there are a variety of ways to represent rotations:

Spherical harmonics

edit
Main article:Spherical harmonics

The groupSO(3) of three-dimensional Euclidean rotations has an infinite-dimensional representation on the Hilbert space

L2(S2)=span{Ym,N+,m},{\displaystyle L^{2}\left(\mathbf {S} ^{2}\right)=\operatorname {span} \left\{Y_{m}^{\ell },\ell \in \mathbb {N} ^{+},-\ell \leq m\leq \ell \right\},} 

whereYm{\displaystyle Y_{m}^{\ell }}  arespherical harmonics. Its elements are square integrable complex-valued functions[nb 5] on the sphere. The inner product on this space is given by

f,g=S2f¯gdΩ=02π0πf¯gsinθdθdϕ.{\displaystyle \langle f,g\rangle =\int _{\mathbf {S} ^{2}}{\overline {f}}g\,d\Omega =\int _{0}^{2\pi }\int _{0}^{\pi }{\overline {f}}g\sin \theta \,d\theta \,d\phi .} H1

Iff is an arbitrary square integrable function defined on the unit sphereS2, then it can be expressed as[19]

|f==1m=m=|YmYm|f,f(θ,ϕ)==1m=m=fmYm(θ,ϕ),{\displaystyle |f\rangle =\sum _{\ell =1}^{\infty }\sum _{m=-\ell }^{m=\ell }\left|Y_{m}^{\ell }\right\rangle \left\langle Y_{m}^{\ell }|f\right\rangle ,\qquad f(\theta ,\phi )=\sum _{\ell =1}^{\infty }\sum _{m=-\ell }^{m=\ell }f_{\ell m}Y_{m}^{\ell }(\theta ,\phi ),} H2

where the expansion coefficients are given by

fm=Ym,f=S2Ym¯fdΩ=02π0πYm¯(θ,ϕ)f(θ,ϕ)sinθdθdϕ.{\displaystyle f_{\ell m}=\left\langle Y_{m}^{\ell },f\right\rangle =\int _{\mathbf {S} ^{2}}{\overline {Y_{m}^{\ell }}}f\,d\Omega =\int _{0}^{2\pi }\int _{0}^{\pi }{\overline {Y_{m}^{\ell }}}(\theta ,\phi )f(\theta ,\phi )\sin \theta \,d\theta \,d\phi .} H3

The Lorentz group action restricts to that ofSO(3) and is expressed as

(Π(R)f)(θ(x),ϕ(x))==1m=m=m=m=Dmm()(R)fmYm(θ(R1x),ϕ(R1x)),RSO(3),xS2.{\displaystyle (\Pi (R)f)(\theta (x),\phi (x))=\sum _{\ell =1}^{\infty }\sum _{m=-\ell }^{m=\ell }\sum _{m'=-\ell }^{m'=\ell }D_{mm'}^{(\ell )}(R)f_{\ell m'}Y_{m}^{\ell }\left(\theta \left(R^{-1}x\right),\phi \left(R^{-1}x\right)\right),\qquad R\in \operatorname {SO} (3),\quad x\in \mathbf {S} ^{2}.} H4

This action is unitary, meaning that

Π(R)f,Π(R)g=f,gf,gS2,RSO(3).{\displaystyle \langle \Pi (R)f,\Pi (R)g\rangle =\langle f,g\rangle \qquad \forall f,g\in \mathbf {S} ^{2},\quad \forall R\in \operatorname {SO} (3).} H5

TheD() can be obtained from theD(m, n) of above usingClebsch–Gordan decomposition, but they are more easily directly expressed as an exponential of an odd-dimensionalsu(2)-representation (the 3-dimensional one is exactly𝖘𝖔(3)).[20][21] In this case the spaceL2(S2) decomposes neatly into an infinite direct sum of irreducible odd finite-dimensional representationsV2i + 1,i = 0, 1, ... according to[22]

L2(S2)=i=0V2i+1i=0span{Ym2i+1}.{\displaystyle L^{2}\left(\mathbf {S} ^{2}\right)=\sum _{i=0}^{\infty }V_{2i+1}\equiv \bigoplus _{i=0}^{\infty }\operatorname {span} \left\{Y_{m}^{2i+1}\right\}.} H6

This is characteristic of infinite-dimensional unitary representations ofSO(3). IfΠ is an infinite-dimensional unitary representation on aseparable[nb 6] Hilbert space, then it decomposes as a direct sum of finite-dimensional unitary representations.[19] Such a representation is thus never irreducible. All irreducible finite-dimensional representations(Π,V) can be made unitary by an appropriate choice of inner product,[19]

f,gUSO(3)Π(R)f,Π(R)gdg=18π202π0π02πΠ(R)f,Π(R)gsinθdϕdθdψ,f,gV,{\displaystyle \langle f,g\rangle _{U}\equiv \int _{\operatorname {SO} (3)}\langle \Pi (R)f,\Pi (R)g\rangle \,dg={\frac {1}{8\pi ^{2}}}\int _{0}^{2\pi }\int _{0}^{\pi }\int _{0}^{2\pi }\langle \Pi (R)f,\Pi (R)g\rangle \sin \theta \,d\phi \,d\theta \,d\psi ,\quad f,g\in V,} 

where the integral is the unique invariant integral overSO(3) normalized to1, here expressed using theEuler angles parametrization. The inner product inside the integral is any inner product onV.

Generalizations

edit

The rotation group generalizes quite naturally ton-dimensionalEuclidean space,Rn{\displaystyle \mathbb {R} ^{n}}  with its standard Euclidean structure. The group of all proper and improper rotations inn dimensions is called theorthogonal group O(n), and the subgroup of proper rotations is called thespecial orthogonal group SO(n), which is aLie group of dimensionn(n − 1)/2.

Inspecial relativity, one works in a 4-dimensional vector space, known asMinkowski space rather than 3-dimensional Euclidean space. Unlike Euclidean space, Minkowski space has an inner product with an indefinitesignature. However, one can still definegeneralized rotations which preserve this inner product. Such generalized rotations are known asLorentz transformations and the group of all such transformations is called theLorentz group.

The rotation group SO(3) can be described as a subgroup ofE+(3), theEuclidean group ofdirect isometries of EuclideanR3.{\displaystyle \mathbb {R} ^{3}.}  This larger group is the group of all motions of arigid body: each of these is a combination of a rotation about an arbitrary axis and a translation, or put differently, a combination of an element of SO(3) and an arbitrary translation.

In general, the rotation group of an object is thesymmetry group within the group of direct isometries; in other words, the intersection of the full symmetry group and the group of direct isometries. Forchiral objects it is the same as the full symmetry group.

See also

edit

Footnotes

edit
  1. '^This is effected by first applying a rotationgθ{\displaystyle g_{\theta }}  throughφ about thez-axis to take thex-axis to the lineL, the intersection between the planesxy andx'y, the latter being the rotatedxy-plane. Then rotate withgθ{\displaystyle g_{\theta }}  throughθ aboutL to obtain the newz-axis from the old one, and finally rotate bygψ{\displaystyle g_{\psi }}  through an angleψ about thenewz-axis, whereψ is the angle betweenL and the newx-axis. In the equation,gθ{\displaystyle g_{\theta }}  andgψ{\displaystyle g_{\psi }}  are expressed in a temporaryrotated basis at each step, which is seen from their simple form. To transform these back to the original basis, observe thatgθ=gϕgθgϕ1.{\displaystyle \mathbf {g} _{\theta }=g_{\phi }g_{\theta }g_{\phi }^{-1}.}  Here boldface means that the rotation is expressed in theoriginal basis. Likewise,
    gψ=gϕgθgϕ1gϕgψ[gϕgθgϕ1gϕ]1.{\displaystyle \mathbf {g} _{\psi }=g_{\phi }g_{\theta }g_{\phi }^{-1}g_{\phi }g_{\psi }\left[g_{\phi }g_{\theta }g_{\phi }^{-1}g_{\phi }\right]^{-1}.} 
    Thus
    gψgθgϕ=gϕgθgϕ1gϕgψ[gϕgθgϕ1gϕ]1gϕgθgϕ1gϕ=gϕgθgψ.{\displaystyle \mathbf {g} _{\psi }\mathbf {g} _{\theta }\mathbf {g} _{\phi }=g_{\phi }g_{\theta }g_{\phi }^{-1}g_{\phi }g_{\psi }\left[g_{\phi }g_{\theta }g_{\phi }^{-1}g_{\phi }\right]^{-1}*g_{\phi }g_{\theta }g_{\phi }^{-1}*g_{\phi }=g_{\phi }g_{\theta }g_{\psi }.} 
  2. ^For an alternative derivation ofso(3){\displaystyle {\mathfrak {so}}(3)} , seeClassical group.
  3. ^Specifically,UJαU=iLα{\displaystyle {\boldsymbol {U}}{\boldsymbol {J}}_{\alpha }{\boldsymbol {U}}^{\dagger }=i{\boldsymbol {L}}_{\alpha }}  for
    U=(i20i2120120i0).{\displaystyle {\boldsymbol {U}}=\left({\begin{array}{ccc}-{\frac {i}{\sqrt {2}}}&0&{\frac {i}{\sqrt {2}}}\\{\frac {1}{\sqrt {2}}}&0&{\frac {1}{\sqrt {2}}}\\0&i&0\\\end{array}}\right).} 
  4. ^For a full proof, seeDerivative of the exponential map. Issues of convergence of this series to the correct element of the Lie algebra are here swept under the carpet. Convergence is guaranteed whenX+Y<log2{\displaystyle \|X\|+\|Y\|<\log 2}  andZ<log2.{\displaystyle \|Z\|<\log 2.}  The series may still converge even if these conditions are not fulfilled. A solution always exists sinceexp is onto in the cases under consideration.
  5. ^The elements ofL2(S2) are actually equivalence classes of functions. two functions are declared equivalent if they differ merely on a set ofmeasure zero. The integral is the Lebesgue integral in order to obtain acomplete inner product space.
  6. ^A Hilbert space is separable if and only if it has a countable basis. All separable Hilbert spaces are isomorphic.

References

edit
  1. ^Jacobson (2009), p. 34, Ex. 14.
  2. ^n × n real matrices are identical to linear transformations ofRn{\displaystyle \mathbb {R} ^{n}}  expressed in itsstandard basis.
  3. ^Coxeter, H. S. M. (1973).Regular polytopes (Third ed.). New York. p. 53.ISBN 0-486-61480-8.{{cite book}}: CS1 maint: location missing publisher (link)
  4. ^Hall 2015 Proposition 1.17
  5. ^Rossmann 2002 p. 95.
  6. ^These expressions were, in fact, seminal in the development of quantum mechanics in the 1930s, cf. Ch III,  § 16, B.L. van der Waerden, 1932/1932
  7. ^Hall 2015 Proposition 3.24
  8. ^Rossmann 2002
  9. ^abEngø 2001
  10. ^Hall 2015 Example 3.27
  11. ^SeeRossmann 2002, theorem 3, section 2.2.
  12. ^Rossmann 2002 Section 1.1.
  13. ^Hall 2003 Theorem 2.27.
  14. ^Shoemake, Ken (1992-01-01), Kirk, DAVID (ed.),"III.6 - Uniform Random Rotations",Graphics Gems III (IBM Version), San Francisco: Morgan Kaufmann, pp. 124–132,ISBN 978-0-12-409673-8, retrieved2022-07-29
  15. ^Hall 2003, Ch. 3;Varadarajan 1984, §2.15
  16. ^Curtright, Fairlie & Zachos 2014 Group elements of SU(2) are expressed in closed form as finite polynomials of the Lie algebra generators, for all definite spin representations of the rotation group.
  17. ^Rodrigues, O. (1840), Des lois géométriques qui régissent les déplacements d'un système solide dans l'espace, et la variation des coordonnées provenant de ses déplacements con- sidérés indépendamment des causes qui peuvent les produire, Journal de Mathématiques Pures et Appliquées de Liouville 5, 380–440.
  18. ^(Goldstein, Poole & Safko 2002, §4.8)
  19. ^abcGelfand, Minlos & Shapiro 1963
  20. ^InQuantum Mechanics – non-relativistic theory byLandau and Lifshitz the lowest orderD are calculated analytically.
  21. ^Curtright, Fairlie & Zachos 2014 A formula forD() valid for all is given.
  22. ^Hall 2003 Section 4.3.5.

Bibliography

edit

[8]ページ先頭

©2009-2025 Movatter.jp