The Matrix Representation of a Three-Dimensional Rotation Revisited

Similar documents
A Primer on Three Vectors

Physics 110. Electricity and Magnetism. Professor Dine. Spring, Handout: Vectors and Tensors: Everything You Need to Know

Index Notation for Vector Calculus

Rotational motion of a rigid body spinning around a rotational axis ˆn;

1.4 LECTURE 4. Tensors and Vector Identities

PHYS 705: Classical Mechanics. Rigid Body Motion Introduction + Math Review

Classical Mechanics Solutions 1.

Vectors. September 2, 2015

Tensors, and differential forms - Lecture 2

Vector calculus. Appendix A. A.1 Definitions. We shall only consider the case of three-dimensional spaces.

Vector and Tensor Calculus

Cartesian Tensors. e 2. e 1. General vector (formal definition to follow) denoted by components

VECTORS, TENSORS AND INDEX NOTATION

1 Vectors and Tensors

NIELINIOWA OPTYKA MOLEKULARNA

The groups SO(3) and SU(2) and their representations

1. Basic Operations Consider two vectors a (1, 4, 6) and b (2, 0, 4), where the components have been expressed in a given orthonormal basis.

,, rectilinear,, spherical,, cylindrical. (6.1)

FoMP: Vectors, Tensors and Fields

Vector analysis and vector identities by means of cartesian tensors

Mathematical Preliminaries

Introduction to Tensor Notation

Classical Mechanics in Hamiltonian Form

A.1 Appendix on Cartesian tensors

Properties of Proper and Improper Rotation Matrices

1.2 Euclidean spacetime: old wine in a new bottle

Three-Dimensional Rotation Matrices

Professor Terje Haukaas University of British Columbia, Vancouver Notation

General tensors. Three definitions of the term V V. q times. A j 1...j p. k 1...k q

L8. Basic concepts of stress and equilibrium

1.3 LECTURE 3. Vector Product

Physics 251 Solution Set 4 Spring 2013

Position and orientation of rigid bodies

A Brief Introduction to Tensors

Electromagnetism HW 1 math review

1. Matrix multiplication and Pauli Matrices: Pauli matrices are the 2 2 matrices. 1 0 i 0. 0 i

1. Rotations in 3D, so(3), and su(2). * version 2.0 *

Tensors - Lecture 4. cos(β) sin(β) sin(β) cos(β) 0

PHY481: Electromagnetism

VECTOR AND TENSOR ALGEBRA

Rigid Bodies. November 5, 2018

1.13 The Levi-Civita Tensor and Hodge Dualisation

Vector analysis. 1 Scalars and vectors. Fields. Coordinate systems 1. 2 The operator The gradient, divergence, curl, and Laplacian...

Homework 7-8 Solutions. Problems

Lecture 19 (Nov. 15, 2017)

Page 52. Lecture 3: Inner Product Spaces Dual Spaces, Dirac Notation, and Adjoints Date Revised: 2008/10/03 Date Given: 2008/10/03

PEAT SEISMOLOGY Lecture 2: Continuum mechanics

PHY481: Electromagnetism

Some elements of vector and tensor analysis and the Dirac δ-function

Chapter 7. Kinematics. 7.1 Tensor fields

Module 4M12: Partial Differential Equations and Variational Methods IndexNotationandVariationalMethods

Physics 221A Fall 1996 Notes 9 Rotations in Ordinary Space

PHYS 4390: GENERAL RELATIVITY NON-COORDINATE BASIS APPROACH

Introduction to Vector Spaces

1 g 1,gg 2 xg 1. 2 g 1,...,gg n xg 1. n g 1 } = {gxg 1,gg 1 x(gg 1 ) 1,gg 2 x(gg 2 ) 1,...,gg n x(gg n ) 1 } = {x,g 1 xg 1. 2,...

Quantum Physics II (8.05) Fall 2004 Assignment 3

Rigid Geometric Transformations

Chapter 1. Describing the Physical World: Vectors & Tensors. 1.1 Vectors

1 Gauss integral theorem for tensors

EN221 - Fall HW # 1 Solutions

Assignment 10. Arfken Show that Stirling s formula is an asymptotic expansion. The remainder term is. B 2n 2n(2n 1) x1 2n.

Introduction to tensors and dyadics

Chapter 2. Linear Algebra. rather simple and learning them will eventually allow us to explain the strange results of

Functional determinants

Review of Linear Algebra

Dr. Tess J. Moon, (office), (fax),

Course 2BA1: Hilary Term 2007 Section 8: Quaternions and Rotations

Physics 218 Polarization sum for massless spin-one particles Winter 2016

Symmetry and Properties of Crystals (MSE638) Stress and Strain Tensor

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition)

Quantum Field Theory III

Group Representations

Covariant electrodynamics

Tensor Analysis in Euclidean Space

2 Tensor Notation. 2.1 Cartesian Tensors

Notes on basis changes and matrix diagonalization

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Chemistry 5.76 Revised February, 1982 NOTES ON MATRIX METHODS

Vectors To begin, let us describe an element of the state space as a point with numerical coordinates, that is x 1. x 2. x =

Week 6: Differential geometry I

Citation for published version (APA): Halbersma, R. S. (2002). Geometry of strings and branes. Groningen: s.n.

III. TRANSFORMATION RELATIONS

Tensor Calculus (à la Speedy Gonzales)

Lecture 10. The Dirac equation. WS2010/11: Introduction to Nuclear and Particle Physics

Introduction to Matrix Algebra

Tensor Calculus. arxiv: v1 [math.ho] 14 Oct Taha Sochi. October 17, 2016

Rigid Geometric Transformations

We are already familiar with the concept of a scalar and vector. are unit vectors in the x and y directions respectively with

Vectors and Matrices Notes.

Introduction to relativistic quantum mechanics

Math review. Math review

Cartesian Tensors. e 2. e 1. General vector (formal definition to follow) denoted by components

There are two things that are particularly nice about the first basis

FORMULA SHEET FOR QUIZ 2 Exam Date: November 8, 2017

Perelman s Dilaton. Annibale Magni (TU-Dortmund) April 26, joint work with M. Caldarelli, G. Catino, Z. Djadly and C.

Matrix Factorization and Analysis

Vectors, metric and the connection

Mathematical Methods wk 2: Linear Operators

ENGI 9420 Lecture Notes 2 - Matrix Algebra Page Matrix operations can render the solution of a linear system much more efficient.

1. Divergence of a product: Given that φ is a scalar field and v a vector field, show that

Solutions for Fundamentals of Continuum Mechanics. John W. Rudnicki

Transcription:

Physics 116A Winter 2010 The Matrix Representation of a Three-Dimensional Rotation Revisited In a handout entitled The Matrix Representation of a Three-Dimensional Rotation, I provided a derivation of the explicit form for most general 3 3 rotation matrix, R(ˆn, θ) that describes the counterclockwise rotation by an angle θ about a fixed axis ˆn. For example, the matrix representation of the counterclockwise rotation by an angle θ about a fixed z-axis is given by [cf. eq. (7.18) on p. 129 of Boas]: cosθ sin θ 0 R(k, θ) sin θ cosθ 0. (1) 0 0 1 The general rotation matrix R(ˆn, θ) satisfies the following relations: R(n, θ + 2πk) = R(n, θ), k = 0, ±1 ± 2..., [R(n, θ)] 1 = R(n, θ) = R( n, θ). (2) Combining these two results, it follows that R(n, 2π θ) = R( n, θ), (3) which implies that any three-dimensional rotation can be described by a counterclockwise rotation by θ about an arbitrary axis ˆn, where 0 θ π. However, if we substitute θ = π in eq. (3), we conclude that R(n, π) = R( n, π), (4) which means that for the special case of θ = π, R(n, π) and R( n, π) represent the same rotation. Finally, if θ = 0, then R(n, 0) = I is the identity operator, independently of the direction of ˆn. In these notes, I will provide a much simpler derivation of the explicit form for R(ˆn, θ), based on the techniques of tensor algebra. 1. A derivation of the Rodriguez formula The matrix elements of R(ˆn, θ) will be denoted by R ij. Since R(ˆn, θ) describes a rotation by an angle θ about an axis ˆn, the formula for R ij that we seek will depend on θ and on the coordinates of ˆn = (n 1, n 2, n 3 ) with respect to a fixed Cartesian coordinate system. Note that since ˆn is a unit vector, it follows that: n 2 1 + n2 2 + n2 3 = 1. (5) 1

Using the techniques of tensor algebra, we can derive the formula for R ij in the following way. We can regard R ij as the components of a second-rank tensor (see Appendix A). Likewise, the n i are components of a vector (equivalently, a first-rank tensor). Two other important quantities for the analysis are the invariant tensors δ ij (the Kronecker delta) and ǫ ijk (the Levi-Civita tensor). If we invoke the covariance of Cartesian tensor equations, then one must be able to express R ij in terms of a second-rank tensor composed of n i, δ ij and ǫ ijk, as there are no other tensors in the problem that could provide a source of indices. Thus, the form of the formula for R ij must be: R ij = aδ ij + bn i n j + cǫ ijk n k, (6) where there is an implicit sum over the index k in the third term of eq. (6). 1 The numbers a, b and c are real scalar quantities. As such, a, b and c are functions of θ, since the rotation angle is the only relevant scalar quantity in this problem. 2 It also follows that ˆn is an axial vector, in which case eq. (6) is covariant with respect to transformations between right-handed and left-handed orthonormal coordinate systems. 3 We now propose to deduce conditions that are satisfied by a, b and c. The first condition is obtained by noting that R(ˆn, θ)ˆn = ˆn. This is clearly true, since R(ˆn, θ), when acting on a vector, rotates the vector around the axis ˆn, whereas any vector parallel to the axis of rotation is invariant under the action of R(ˆn, θ). In terms of components R ij n j = n i. (7) To determine the consequence of this equation, we insert eq. (6) into eq. (7) and make use of eq. (5). Noting that 4 δ ij n j = n i, n j n j = 1 ǫ ijk n j n k = 0, (8) 1 We follow the Einstein summation convention in these notes. That is, there is an implicit sum over any pair of repeated indices in the present and all subsequent formulae. 2 One can also construct a scalar by taking the dot product of ˆn ˆn, but this quantity is equal to 1 [cf. eq. (5)], since ˆn is a unit vector. Thus, it does not add anything new. 3 Under inversion of the coordinate system, θ θ and ˆn ˆn. However, since 0 θ π (by convention), we must then use eq. (2) to flip the signs of both θ and ˆn to represent the rotation R(ˆn, θ) in the new coordinate system. Hence, the signs of θ and ˆn effectively do not change under the inversion of the coordinate system. That is, θ is a scalar and ˆn is an axial (or pseudo-) vector. 4 In the third equation of eq. (8), there is an implicit sum over j and k. Since ǫ ijk = ǫ jik, when the sum ǫ ijk n j n k is carried out, we find that for every positive term, there is an identical negative term to cancel it. The total sum is therefore equal to zero. This is an example of a very general rule. Namely, one always finds that the product of two tensor quantities, one symmetric under the interchange of a pair of summed indices and one antisymmetric under the interchange of a pair of summed indices, is equal to zero when summed over the two indices. In the present case, n j n k is symmetric under the interchange of j and k, whereas ǫ ijk is antisymmetric under the interchange of j and k. Hence their product, summed over j and k, is equal to zero. 2

it follows immediately that n i (a + b) = n i. Hence, a + b = 1. (9) Since the formula for R ij given by eq. (6) must be completely general, it must hold for any special case. In particular, consider the case where ˆn = k. In this case, eqs. (1) and (6) yields: R(k, θ) 11 = cosθ = a, R(k, θ) 12 = sin θ = c ǫ 123 n 3 = c. (10) Using eqs. (9) and (10) we conclude that, a = cos θ, b = 1 cosθ, c = sin θ. (11) Inserting these results into eq. (6) yields the Rodriguez formula: R ij (ˆn, θ) = cosθ δ ij + (1 cosθ)n i n j sin θ ǫ ijk n k (12) We can write the above quantity in 3 3 matrix form, although eq. (12) is more compact and convenient. For completeness, here is the explicit form for the general 3 3 rotation matrix: cosθ + n 2 1(1 cosθ) n 1 n 2 (1 cos θ) n 3 sin θ n 1 n 3 (1 cos θ) + n 2 sin θ R(ˆn, θ) = n 1 n 2 (1 cosθ) + n 3 sin θ cosθ + n 2 2 (1 cosθ) n 2n 3 (1 cosθ) n 1 sin θ, n 1 n 3 (1 cosθ) n 2 sin θ n 2 n 3 (1 cosθ) + n 1 sin θ cosθ + n 2 3(1 cosθ) (13) where n 2 1 + n2 2 + n2 3 = 1. Thus, we have reproduced the explicit form for the general 3 3 rotation matrix, which was derived in the previous handout, The Matrix Representation of a Three-Dimensional Rotation. 2. Determining the rotation axis and the rotation angle In Section 3 of the previous handout, The Matrix Representation of a Three- Dimensional Rotation, I presented an algorithm for obtaining the direction of the rotation axis ˆn and the rotation angle θ if we are given an arbitrary 3 3 rotation matrix R(ˆn, θ). With some tensor algebra manipulations involving the Levi-Civita tensor, we can use eq. (12) to quickly obtain the desired results. First, we compute the trace of R(ˆn, θ). In particular, using eq. (12) it follows that: Tr R(ˆn, θ) = R ii = 1 + 2 cosθ (14) In deriving this result, we used the fact that δ ii = Tr I = 3 (since the indices run over i = 1, 2, 3 in three-dimensional space) and ǫ iik = 0 (the latter is a consequence of the fact that the Levi-Civita tensor is totally antisymmetric under the interchange of any two indices). By convention, 0 θ π, which implies that sin θ 0. Thus, cos θ = 1 2 (R ii 1) and sin θ = (1 cos 2 θ) 1/2 = 1 2 (3 Rii )(1 + R ii ) (15) 3

where cosθ is determined from eq. (14). All that remains is to determine the axis of rotation ˆn. Let us multiply eq. (12) by ǫ ijm and sum over i and j. Noting that 5 it follows that ǫ ijm δ ij = ǫ ijm n i n j = 0, ǫ ijk ǫ ijm = 2δ km, (16) 2n m sin θ = R ij ǫ ijm. (17) If R is a symmetric matrix (i.e. R ij = R ji ), then R ij ǫ ijm = 0 automatically (since ǫ ijk is antisymmetric under the interchange of the indices i and j). In this case sin θ = 0 and we must seek other means to determine ˆn. If sin θ 0, then one can divide both sides of eq. (17) by sin θ. Using eq. (15), we obtain: 6 n m = R ijǫ ijm 2 sin θ = R ij ǫ ijm, sin θ 0 (18) (3 Rii )(1 + R ii ) More explicitly, ) 1 ˆn = (R 32 R 23, R 13 R 31, R 21 R 12, R ii 1, 3. (19) (3 Rii )(1 + R ii ) In Appendix B, we verify that ˆn as given by eq. (18) is a vector of unit length [as required by eq. (5)]. The overall sign of ˆn is fixed by eq. (18) due to our convention in which sin θ 0. If we multiply eq. (17) by n m and sum over m, then sin θ = 1 2 ǫ ijmr ij n m, (20) after using n m n m = 1. This provides an additional check on the determination of the rotation angle. As noted above, if R is a symmetric matrix (i.e. R ij = R ji ), then sin θ = 0 and ˆn cannot be determined from eq. (18). In this case, eq. (14) determines whether cosθ = +1 or cosθ = 1. If cosθ = +1, then R ij = δ ij and the axis ˆn is undefined. If cosθ = 1, then according to eq. (12), R ij = 2n i n j δ ij, which determines the direction of ˆn up to an overall sign. That is, ˆn is undetermined if θ = 0, ) 1 ˆn = (ǫ 1 (1 + R 1 2 11), ǫ 2 (1 + R 1 2 22), ǫ 3 (1 + R 2 33), if θ = π, (21) 5 In regards to the first equation of eq. (16), see the footnote 4. The second equation of eq. (16) is given in eq. (5.8) on p. 511 of Boas. 6 If sin θ = 0, both the numerator and denominator of eq. (18) vanish. We will show below [cf. eq. (21)] that ˆn is undefined if R ii = 3, corresponding to the case of R(ˆn, 0) = I. When R ii = 1, corresponding to the case of R(ˆn, π), ˆn can be determined directly from eq. (12). 4

where the individual signs ǫ i = ±1 are determined up to an overall sign via 7 ǫ i ǫ j = R ij (1 + Rii )(1 + R jj ), for fixed i j, R ii 1, R jj 1. (22) The ambiguity of the overall sign of ˆn sign is not significant, since R(ˆn, π) and R( ˆn, π) represent the same rotation [cf. eq. (4)]. One slightly inconvenient feature of the above analysis is that the case of R ii = 1 (corresponding to θ = π) requires a separate treatment in order to determine ˆn. Moreover, for values of θ very close to π, the numerator and denominator of eq. (19) are very small, so that a very precise numerical evaluation of both the numerator and denominator is required to accurately determine the direction of ˆn. Thus, we briefly mention another approach for determining ˆn that can be employed for all possible values of R ii (except for R ii = 3 corresponding to the identity rotation, where ˆn is not defined). This approach is based directly on the Rodriguez formula [eq. (12)]. Define the matrix S = R + R T + (1 R ii )I. (23) Then, eq. (12) yields S jk = 2(1 cosθ)n j n k = (3 R ii )n j n k. Hence, 8 n j n k = S jk 3 R ii, R ii 3 (24) Note that for θ close to π (which corresponds to R ii 1), neither the numerator nor the denominator of eq. (24) is particularly small, and the direction of ˆn can be determined numerically without significant roundoff error. To determine ˆn up to an overall sign, we simply set j = k (no sum) in eq. (24), which fixes the value of n 2 j. If sin θ 0, the overall sign of ˆn is fixed by eq. (17). If sin θ = 0 there are two cases. For θ = 0 (corresponding to the identity rotation), the rotation axis ˆn is undefined. For θ = π, the ambiguity in the overall sign of ˆn is immaterial, in light of eq. (4). Thus, we have achieved our goal. Eqs. (15), (19) and (21) [or equivalently, eqs. (15), (17) and (24)] provide a simple algorithm for determining the rotation axis ˆn and the rotation angle θ for any rotation matrix R(ˆn, θ) I. 7 If R ii = 1 [no sum over i], then n i = 0, in which case the corresponding ǫ i is not well-defined. 8 Eq. (23) yields TrS S ii = 3 R ii. One can then use eq. (24) to verify that ˆn is a unit vector. 5

Appendix A: Matrix elements of matrices correspond to the components of second rank tensors In the class handout entitled Coordinates, matrix elements and changes of basis, we examined how the matrix elements of linear operators change under a change of basis. Consider the matrix elements of a linear operator with respect to two different orthonormal bases, B = {ê 1, ê 2, ê 3 } and B = {ê 1, ê 2, ê 3}. Then, using the Einstein summation convention, ê j = P ijê i, where P is an orthogonal matrix. Given any linear operator A with matrix elements a ij with respect to the basis B, the matrix elements a ij with respect to the basis B are given by a kl = (P 1 ) ki a ij P jl = P ik a ij P jl, where we have used the fact that P 1 = P T in the second step above. Finally, identifying P = R 1, where R is also an orthogonal matrix, it follows that a kl = R kir lj a ij, which we recognize as the transformation law for the components of a second rank Cartesian tensor. Appendix B: Verifying that ˆn as determined from eq. (18) is a unit vector We first need some preliminary results. Using the results from the handout entitled The Characteristic Polynomial, the characteristic equation of an arbitrary 3 3 matrix R is given by: p(λ) = [ λ 3 λ 2 Tr R + c 2 λ 2 det R ], where c 2 = 1 2 [ (Tr R) 2 Tr(R 2 ) ]. For a special orthogonal matrix, det R = 1. Hence, p(λ) = [ λ 3 λ 2 Tr R + 1 2 λ [ (Tr R) 2 Tr(R 2 ) ] 1 ]. We now employ the Cayley-Hamilton theorem, which states that a matrix satisfies its own characteristic equation, i.e. p(r) = 0. Hence, R 3 R 2 Tr R + 1 2 R [ (Tr R) 2 Tr(R 2 ) ] I = 0. Multiplying the above equation by R 1, and using the fact that R 1 = R T for an orthogonal matrix, R 2 R Tr R + 1 2 I [ (Tr R) 2 Tr(R 2 ) ] R T = 0. 6

Finally, we take the trace of the above equation. Using Tr(R T ) = Tr R, we can solve for Tr(R 2 ). Using Tr I = 3, the end result is given by: Tr(R 2 ) = (Tr R) 2 2 TrR, (25) which is satisfied by all 3 3 special orthogonal matrices. We now verify that ˆn as determined from eq. (18) is a unit vector. For convenience, we repeat eq. (18) here: n m = 1 R ij ǫ ijm 2 sin θ = R ij ǫ ijm (3 Rii )(1 + R ii ), sin θ 0, where R ii Tr R. We evaluate ˆn ˆn = n m n m as follows: n m n m = R ijǫ ijm R kl ǫ klm (3 R ii )(1 + R ii ) = R ijr kl (δ ik δ jl δ il δ jk ) (3 R ii )(1 + R ii ) The numerator of the above expression is equal to: R ij R ij R ij R ji = Tr(R T R) Tr(R 2 ) = Tr I Tr(R 2 ) = R ijr ij R ij R ji (3 R ii )(1 + R ii ). = 3 Tr(R 2 ) = 3 (TrR) 2 + 2 TrR = (3 R ii )(1 + R ii ), (26) after using eq. (25) for Tr(R 2 ). Hence, employing eq. (26) yields and the proof is complete. ˆn ˆn = n m n m = R ijr ij R ij R ji (3 R ii )(1 + R ii ) = 1, 7