<<

KPOP E-2020-04

Bra-Ket Representation of the Inertia

U-Rae Kim, Dohyun Kim, and Jungil Lee∗ KPOPE Collaboration, Department of , Korea University, Seoul 02841, Korea (Dated: June 18, 2021) Abstract We employ Dirac’s bra-ket notation to define the inertia tensor that is independent of the choice of bases or . The principal axes and the corresponding principal values for the elliptic plate are determined only based on the . By making use of a general symmetric , we develop a method of diagonalization that is convenient and intuitive in determining the eigenvector. We demonstrate that the bra-ket approach greatly simplifies the computation of the inertia tensor with an example of an N-dimensional ellipsoid. The exploitation of the bra-ket notation to compute the inertia tensor in classical mechanics should provide undergraduate students with a strong background necessary to deal with abstract quantum mechanical problems.

PACS numbers: 01.40.Fk, 01.55.+b, 45.20.dc, 45.40.Bb

Keywords: Classical mechanics, Inertia tensor, Bra-ket notation, Diagonalization, Hyperellipsoid arXiv:2012.13347v1 [physics.class-ph] 15 Dec 2020

∗Electronic address: [email protected]; Director of the Korea Pragmatist Organization for Physics Educa-

tion (KPOP E)

1 I. INTRODUCTION

The inertia tensor is one of the essential ingredients in classical mechanics with which one can investigate the rotational properties of rigid-body motion [1]. The symmetric nature of the rank-2 guarantees that it is described by three fundamental parameters called the principal moments of inertia Ii, each of which is the along a principal axis. The principal axes are orthogonal so that they construct a natural choice of the Cartesian coordinate system. Additional three parameters of the can be allocated to the from the natural coordinate system of the principal axes to an arbitrary coordinate system. Thus, the tensor is one of the best tools with which students can train on how to deal with the eigenvalue problem of a Hermitian . The underlying mathematical structure of this problem is closely related to the modes. It can further be extended to the Hilbert space in the continuum limit [2]. Hence, we believe that teaching inertia is of great importance in giving students a strong background for their future study of . Many pedagogical materials involving the general and the inertia tensors are available [3– 8], including undergraduate level textbooks [1, 2]. Most of them deal with the inertia tensor in explicit matrix representations. In this paper, we employ Dirac’s bra-ket notation [9–11] to define the inertia tensor operator, from which one can project out the corresponding inertia tensor elements by applying a bra and a ket for the unit vectors qˆi along the principal axes. While the completeness relation reflects the homogeneousness and isotropy of the 1 = |qˆiihqˆi|, the inertia tensor operator acquires the anisotropy of the mass distribution so that each principal axis is distinct in general: I = Ii|qˆiihqˆi|. The coordinate dependence of the inertia tensor of a system with the unit basis vector eˆi can be read off directly from the operator by finding the projections: I = (Iij) = (heˆi|I|eˆji). The operator method is independent of the coordinate system. Thus, the computation of the matrix element is minimal because one needs to compute only three fundamental integrals Ii that are the principal moments. An extension to the covariant approach as in [12] is straightforward because of the coordinate independence of this formulation. This paper is organized as follows: In Sec. II, we list the definitions of the quantities involving the bra-ket notation and the inertia tensor operator. A summary of the relationship between the principal axes and the coordinate dependence of the inertia tensor is given in

2 Sec. III. As applications of the approach developed in the preceding sections, we compute the inertia tensor for an ellipsoid in the three- and N-dimensional Euclidean spaces in Sec. IV and we conclude in Sec. V. Appendices are attached to list the basis operators for the rank-2 Cartesian tensors in two and three .

II. DEFINITIONS

A. Bra-Ket Notation

By employing Dirac’s bra-ket notation [9–11], one can express the scalar product of any two Euclidean vectors U and V in N dimensions as

U · V = hU|Vi = hV|Ui. (1)

Let us consider unit basis vectors eˆi in the Cartesian coordinate system that satisfy the orthonormal condition and the completeness relation

heˆj|eˆki = δjk, (2)

|eˆiiheˆi| = 1, (3)

where δjk is the and i is summed over 1, ··· , N. In the remainder of this paper, any repeated vector indices are assumed to be summed over unless specified. Any vector can be expanded in terms of the bases as

|Vi = Vi|eˆii, (4)

where the Cartesian component Vi of V can be read off as

Vi = heˆi|Vi. (5)

The identity operator 1 is invariant under rotation. Thus, the corresponding matrix repre- sentation 1 is independent of the coordinate system:

1 = (heˆi|1|eˆji) = (heˆi|eˆji) = (δij). (6)

3 B. Inertia Tensor Operator

The ij element of the inertia tensor of a is a symmetric rank-2 Cartesian tensor: Z 3 2 Iij = d x ρ(x)[δijx − xixj], (7) where ρ(x) is the mass density at point x and the integration is over the space. The rotational properties of the inertia tensor is completely determined by the scalar field ρ(x). By employing Dirac’s bra-ket notation, we can express the inertia tensor operator as Z I = d3xρ(x)hx|xi1 − |xihx| . (8)

The matrix representation of the inertia tensor I = (Iij) in the coordinate system with the basis vectors eˆi’s can be expressed as

Iij = heˆi|I|eˆji. (9)

Note that the quantities like Vi in Eq. (4) or Iij in Eq. (7) depend on the choice of basis vectors or coordinate system. However, the operator representation as in Eq. (8) is independent of the coordinate system unless it is rotationally invariant as in Eq. (3).

III. GEOMETRY AND PRINCIPAL AXES

We consider an elliptic plate that has the simplest geometry without having rotational symmetry in two dimensions. The principal axes and the corresponding principal values for the rigid body are first considered, not based on the inertia tensor but based on the geometry only, through the parametric equation. The operator for the geometry of the rigid body is extended to the inertia tensor operator.

A. Principal Axes

The parametric curve of an ellipse on the xy plane is x2 y2 + = 1, (10) a2 b2 where a and b are the semi-major and semi-minor axes that are along the x and y axes, respectively. An important feature of this expression is that the left side is the sum of squared

4 numbers. We call the coordinate axes of the system the principal axes with which a physical observable can be expressed as the sum of squared numbers. In this case, those principal axes consist of the major and the minor axes. Thus, any expression for an observable acquires the most compact form. This is a reflection of the symmetries of a physical observable.

We define x1 ≡ x, x2 ≡ y, and the vectors qˆi’s along the principal axes to find that

|xi = xi|qˆii. (11)

The parametric equation (10) can be represented by

hx|SP|xi = 1, (12)

where SP is the tensor operator defined by

SP = αP|qˆ1ihqˆ1| + βP|qˆ2ihqˆ2|. (13)

The operator SP is a special case of γ = 0 in a general symmetric tensor operator in Eq. (A1). The subscript P indicates that the operator is written in the frame whose coordinate axes are principal axes. Here, the coefficients 1 α = , (14a) P a2 1 β = , (14b) P b2

are the principal values for the operator SP corresponding to principal axes qˆ1 and qˆ2, respectively. The operator for the inertia tensor for a two-dimensional rigid body must be of the form

I = I1|qˆ1ihqˆ1| + I2|qˆ2ihqˆ2|. (15)

This is true for any two-dimensional rigid body as well as the elliptic plate that we consider in this section. The reason is that we can always construct an equivalent elliptic plate that has the same principal axes with the same principal values.

B. Non-principal Axes

We shall find that the operator SP in Eq. (13) acquires additional contributions in a frame of reference whose axes are not principal. We consider another set of orthonormal

5 FIG. 1: An elliptic plate with semi-major axis a and semi-minor axis b. The principal axes qˆi’s

of the plate are related to eˆi’s under rotation by an angle θ.

basis vectors eˆi’s that are related to the principal axes under rotation as shown in Fig. 1:

|qˆ1i = cos θ|eˆ1i + sin θ|eˆ2i, (16)

|qˆ2i = − sin θ|eˆ1i + cos θ|eˆ2i. (17)

The corresponding directional cosines are given by

heˆ1|qˆ1i = cos θ, heˆ2|qˆ1i = sin θ,

heˆ1|qˆ2i = − sin θ, heˆ2|qˆ2i = cos θ.

By multiplying the identity operators in Eq. (3) to both sides of Eq. (13), we can express

the operator SP in terms of the basis kets |eˆii’s as

SP = 1 SP1   = αPheˆi|qˆ1ihqˆ1|eˆji + βPheˆi|qˆ2ihqˆ2|eˆji |eˆiiheˆj|

1 1 = 2 (αP + βP)1 + 2 (αP − βP)(S1 cos 2θ + S2 sin 2θ)

≡ s01 + s1S1 + s2S2, (18)

where the operators Si’s are defined in Eqs. (A5), (A6), and (A7) and

1 s0 = 2 (αP + βP), (19a) 1 s1 = 2 (αP − βP) cos 2θ, (19b) 1 s2 = 2 (αP − βP) sin 2θ. (19c)

The off-diagonal contribution emerges through the contribution proportional to S2. Note that the parametrization in Eq. (18) is applicable to any real symmetric rank-2 tensor operator.

6 2 2 Additionally, s0 and s1 + s2 are invariant separately: q 2 2 1 1 s1 + s2 = 2 |αP − βP| = 2 (αP − βP), (20) 2 2 2 1 2 2 s0 + s1 + s2 = 2 (αP + βP), (21) where  is the sign of αP − βP:

 = sgn[αP − βP]. (22)

The angle θ can be determined as s cos 2θ = 1 , (23) p 2 2 s1 + s2 s sin 2θ = 2 , (24) p 2 2 s1 + s2 1 s θ = arctan 2 . (25) 2 s1 The principal values can be expressed as q 2 2 αP = s0 +  s1 + s2, (26) q 2 2 βP = s0 −  s1 + s2, (27) which correspond, respectively, to the principal axes qˆ1 and qˆ2.

C. Diagonalization

We determine the angle to diagonalize a real symmetric rank-2 tensor operator S in Eq. (A1),

S = s01 + s1S1 + s2S2, (28) to reach the form in Eq. (13). In most textbooks, the readers are advised to solve the eigenvalue problem to find the principal axes. Although such an approach is mathematically straightforward, it does not tend to provide students with physical intuitions. Thus, one must investigate the solution further to interpret the physical meaning even after obtaining the result. Here, we rotate the orthonormal basis vectors from {eˆ1, eˆ2} to {qˆ1, qˆ2} as shown in Fig. 1 as

|eˆ1i = cos θ|qˆ1i − sin θ|qˆ2i, (29)

|eˆ2i = sin θ|qˆ1i + cos θ|qˆ2i. (30)

7 Substituting these into Eq. (28), we obtain

S = (s0 + s1 cos 2θ + s2 sin 2θ)|qˆ1ihqˆ1|

+(s0 − s1 cos 2θ − s2 sin 2θ)|qˆ2ihqˆ2|

+(−s1 sin 2θ + s2 cos 2θ)(|qˆ1ihqˆ2| + |qˆ2ihqˆ1|). (31)

The operator S reaches the principal form in which the crossing terms |qˆ1ihqˆ2| and

|qˆ2ihqˆ1| vanish:

S = αP|qˆ1ihqˆ1| + βP|qˆ2ihqˆ2|, (32) where αP and βP are principal values in Eq. (14). The vanishing condition is satisfied when

s1 sin 2θP = s2 cos 2θP. (33)

The angle θP can always be chosen as

1 s2 θP = arctan , 0 ≤ θP ≤ π/2. (34) 2 s1 A convenient choice of the solution set is 1 (cos 2θP, sin 2θP) = (s1, s2). (35) p 2 2 s1 + s2 The principal values are determined as

αP = s0 + s1 cos 2θP + s2 sin 2θP q 2 2 = s0 + s1 + s2, (36)

βP = s0 − s1 cos 2θP − s2 sin 2θP q 2 2 = s0 − s1 + s2, (37) where we have made use of the solution set in Eq. (35). By making use of the trigonometric identity cos2 θ − sin2 θ = cos 2θ and applying the choice in Eq. (34), we can always determine cos θP and sin θP uniquely as

q 1 cos θP = 2 (1 + cos 2θP) v u " # u1 s1 = t 1 + , (38a) 2 p 2 2 s1 + s2

q 1 sin θP = 2 (1 − cos 2θP) v u " # u1 s1 = t 1 − . (38b) 2 p 2 2 s1 + s2

8 The corresponding principal axes are

|qˆ1i = cos θP|eˆ1i + sin θP|eˆ2i, (39)

|qˆ2i = − sin θP|eˆ1i + cos θP|eˆ2i. (40)

Our determination of the that diagonalizes the symmetric tensor oper- ator is manifest. This approach is particularly convenient and intuitive in determining the eigenvector.

IV. INERTIA TENSOR OF AN ELLIPSOID

In Sec. III, we found that the parametric equation for the ellipse can be expressed mini- mally by choosing the principal axes as the coordinate axes. The same principal axes can be used to express the inertia tensor that shares the principal axes because it is a fundamen- tal property of symmetric tensor operators. In this section, we compute the inertia tensor operators for uniform ellipsoids in the three- and N-dimensional Euclidean spaces. Because the inertia tensor operator for the elliptic plate can be obtained from the three-dimensional counterpart, we omit that calculation.

A. Ellipsoid in Three Dimensions

As an application, we compute the inertia tensor for a uniform ellipsoid of mass M in three-dimensional Euclidean space. The parametric equation (10) can be generalized to P3 2 2 i=1 xi /ai = 1. Then the mass distribution is 3M ρ(x, q1, q2, q3) = Θ[1 − hx|E|xi], (41) 4πa1a2a3 where Θ is the Heaviside step and E is a symmetric tensor operator 1 E = 2 |qˆiihqˆi|. (42) ai

Here, qˆi are the orthonormal basis vectors along the principal axes of the ellipsoid. Because a real symmetric rank-2 tensor operator in three dimensions can be expressed as Eq. (B1), the inertia tensor operator must be expanded as Z 3   I = d x ρ(x, q1, q2, q3) hx|xi1 − |xihx| X = αi|qˆiihqˆi| + γijSij, (43) i

9 where Sij = |qˆiihqˆj| + |qˆjihqˆi|. The scalar factors in Eq. (43) can be read off as

αi = hqˆi|I|qˆii, (44)

γij = hqˆi|I|qˆji. (45)

Therefore, the matrix element Iij = heˆi|I|eˆji of the inertia tensor I can be determined as X Iij = αkqkiqkj + γk` (qkiq`j + q`iqkj) , (46) k<` where qki ≡ hqˆk|eˆii = qˆk · eˆi.

Let us consider the scalar integrals for αi and γij: Z 3M 3 2 αi = d x (hx|xi − hx|qˆii ), 4πa1a2a3 V Z 3M 3 γij = − d xhqˆi|xihx|qˆji, (47) 4πa1a2a3 V where Z Z d3x ≡ d3x Θ[1 − hx|E|xi]. (48) V Here, we denote V by the region inside the ellipsoid. If we define

hqˆi|xi y = qˆi, (49) ai then the region V is parametrized by

hy|yi ≤ 1. (50)

In addition, we find that Z Z 3 3 d x = a1a2a3 d y, (51a) V V 2 2 hx|xi = hx|qˆiihqˆi|xi = akhy|qˆki , (51b)

2 2 2 hx|qˆii = ai hy|qˆii , (51c)

where i in the last line is not summed over. Making a change of variable in Eq. (47) by using the identities in Eq. (51) and carrying out the integration, we obtain Z 3M 3 2 αi = d x (hx|xi − hx|qˆii ) 4πa1a2a3 V Z 3M 2 3 2 = ak(1 − δki) d y hy|qˆki , 4π V Z 3M 3 γij = − d xhqˆi|xihx|qˆji 4πa1a2a3 V = 0, (52)

10 where γij = 0 because the integrand is an odd function of both hx|qˆii and hx|qˆji while the region V is even in those variables. Note that the factor (1 − δki) in αi is the projection op- erator for the direction perpendicular to qˆi. The isotropy of the Euclidean space guarantees that Z Z 3 2 1 3 d y hy|qˆki = d y hy|yi V 3 V 4π Z 1 = dr r4 3 0 4π = . (53) 15

As a result, the principal moments are determined as

1 α = M(a2 + a2), 1 5 2 3 1 α = M(a2 + a2), 2 5 3 1 1 α = M(a2 + a2). (54) 3 5 1 2

B. Ellipsoid in N Dimensions

As a pedagogical example, we evaluate the inertia tensor for a uniform ellipsoid of mass M in N-dimensional Euclidean space. The parametric equation of the ellipsoid in N dimensions is N X x2 i = 1. (55) a2 i=1 i The of the ellipsoid is given by Z N VN = d x Θ[1 − hx|EN |xi], (56)

where EN is a symmetric tensor operator

N N X X 1 E = |q ihq | = |qˆ ihqˆ |. (57) N i i a2 i i i=1 i=1 i

Note that qˆi are the orthonormal basis vectors along the principal axes of the ellipsoid. Introducing the rescaled vector y which is the N-dimensional generalization of Eq. (49), we

11 find that N ! Z Y N 2 VN = ai d y Θ[1 − y ] i=1 N ! Z 1 Y N−1 = ai ΩN dr r , i=1 0 N ΩN Y = a , (58) N i i=1 where ΩN is the N-dimensional solid angle [12] 1 N 2π 2 ΩN = 1 . (59) Γ( 2 N) The mass distribution is NM ρ(x, q , ··· , q ) = Θ[1 − hx|E |xi]. (60) 1 N QN N ΩN i=1 ai Because the inertia tensor operator is symmetric, it must be expanded as Z N   I = d x ρ(x, q1, ··· , qN ) hx|xi1 − |xihx|

N X X = αi|qˆiihqˆi| + γijSij, (61) i=1 i

where Sij = |qˆiihqˆj| + |qˆjihqˆi|. The scalar factors in Eq. (61) can be determined as

αi = hqˆi|I|qˆii, (62)

γij = hqˆi|I|qˆji. (63)

Therefore, the matrix element Iij = heˆi|I|eˆji of the inertia tensor I can be computed as N X X Iij = αkqkiqkj + γk` (qkiq`jq`iqkj) , (64) k=1 k<`

where qki ≡ hqˆk|eˆii = qˆk · eˆi.

We evaluate the scalar integrals for αi and γij as NM Z α = dN x (hx|xi − hx|qˆ i2) i QN i ΩN i=1 ai V Z NM 2 N 2 = ak(1 − δki) d y hy|qˆki , (65) ΩN V NM Z γ = − dN xhqˆ |xihx|qˆ i ij QN i j ΩN i=1 ai V = 0, (66)

12 where γij = 0 because both hx|qˆii and hx|qˆji in the integrand are odd while the region V is even in those variables. We notice that (1 − δki) is the projection operator for the direction perpendicular to qˆi. By making use of the isotropy of the Euclidean space, we obtain Z Z N 2 1 N d y hy|qˆki = d y hy|yi V N V Ω Z 1 = N dr rN+1 N 0 Ω = N . (67) N(N + 2) Therefore, M X α = a2, (68) i N + 2 k k6=i which reproduces the result in Eq. (54) when N = 3.

V. DISCUSSION

In this paper, we have employed Dirac’s bra-ket notation [9–11] to define the inertia tensor operator I in Eq. (8). This operator is independent of the choice of bases or coordinate system. The matrix elements Iij of the inertia tensor can be read off by applying the bra and the ket for the unit basis vectors eˆi as Iij = heˆi|I|eˆji. The computation of the matrix element is minimal when the principal axes are taken to be the bases because one needs only to compute the three fundamental scalar integrals Ii that are the principal moments. We have considered an elliptic plate, which has the simplest geometry without having rotational symmetry in two dimensions. The principal axes and the corresponding principal values for the elliptic plate have been determined only based on the geometry through the introduction of the symmetric operator that satisfies the parametric equation of the ellipse. It is worthwhile to mention that the inertia tensor operator of any planar lamina can always be represented by the symmetric tensor of the elliptic plate that has the same principal values. By making use of a general symmetric tensor operator in Eq. (A1), we have constructed the method of diagonalization without solving an eigenvalue problem. This approach is particularly convenient and intuitive in determining the eigenvector. As applications, we have determined the inertia tensor operators for uniform ellipsoids in three and N dimensions. The bra-ket approach allows one to express the tensor operator in a compact form, and one has to evaluate only the minimal number of the scalar integrals, the

13 principal moments of inertia. We have demonstrated that the bra-ket notation interplayed with various rescaling of the state kets greatly simplifies the evaluation of the scalar integrals corresponding to the principal moments of the N-dimensional ellipsoid. The bra-ket approach, which is independent of the coordinate choice is naturally extended to the covariant approach that is given, for example, in [12]. The exploitation of the bra-ket notation to compute the inertia tensor in classical mechanics should provide undergraduate students with a strong background that may result in a smooth transition to attack abstract quantum mechanical problems.

Acknowledgments

As members of the Korea Pragmatist Organization for Physics Education (KPOP E), the authors thank the remaining members of KPOP E for useful discussions. This work is supported in part by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) under Contract Nos. NRF-2020R1A2C3009918 (U-R.K. and J.L.), NRF-2017R1E1A1A01074699 (J.L.), and NRF-2019R1A6A3A01096460 (U-R.K.), and the BK21+ program at Korea University, Initiative for Creative and Independent Scientists.

Appendix A: Basis Tensor Operator in Two Dimensions

In two dimensions, at most two linearly independent vectors are available. By making use of the Gram-Schmidt orthogonalization procedure, we can always express the operator in terms of two orthogonal vectors a and b. We set eˆ1 and eˆ2 to be the orthonormal basis vectors along these directions. The real symmetric tensor operator must be of the form

S = α|eˆ1iheˆ1| + β|eˆ2iheˆ2| + γ(|eˆ1iheˆ2| + |eˆ2iheˆ1|)

= siSi, (A1)

where si is a scalar 1 s = (α + β), (A2) 0 2 1 s = (α − β), (A3) 1 2

s2 = γ. (A4)

14 Also, the operators Si are defined by

S0 = |eˆ1iheˆ1| + |eˆ2iheˆ2| = 1, (A5)

S1 = |eˆ1iheˆ1| − |eˆ2iheˆ2|, (A6)

S2 = |eˆ1iheˆ2| + |eˆ2iheˆ1|. (A7)

From these definitions, we find that

1 |eˆ iheˆ | = (S + S ), (A8) 1 1 2 0 1 1 |eˆ iheˆ | = (S − S ). (A9) 2 2 2 0 1

These basis tensor operators Si’s have a common feature:

2 Si = 1, (A10) where i = 0, 1, 2 and the vector index i is not summed over.

Their matrix representations Sk ≡ (heˆi|Sk|eˆji) are   1 0 S 0 =   , (A11) 0 1   1 0 S 1 =   , (A12) 0 −1   0 1 S 2 =   . (A13) 1 0

Note that only S0 is traceful, and that the remaining matrices S1 and S2 are both symmetric and traceless. The last two matrices are identical to the Pauli sigma matrices: S1 = σ3 and

S2 = σ1.

Appendix B: Basis Tensor Operator in Three Dimensions

In three dimensions, at most three linearly independent vectors are available. By making use of the Gram-Schmidt orthogonalization procedure, we can always express the operator in terms of three orthogonal vectors eˆ1, eˆ2, and eˆ3. The real symmetric tensor operator

15 must be of the form

S = α|eˆ1iheˆ1| + β|eˆ2iheˆ2| + γ|eˆ3iheˆ3| + µ(|eˆ1iheˆ2| + |eˆ2iheˆ1|)

+ν(|eˆ1iheˆ3| + |eˆ3iheˆ1|) + λ(|eˆ2iheˆ3| + |eˆ3iheˆ2|)

= siSi, (B1)

where si is a scalar 1 s = (α + β + γ), (B2) 0 3 1 s = (α − β), (B3) 1 2 1 s2 = √ (α + β − 2γ), (B4) 2 3 s3 = µ, (B5)

s4 = ν, (B6)

s5 = λ, (B7) and

S0 = |eˆ1iheˆ1| + |eˆ2iheˆ2| + |eˆ3iheˆ3| = 1, (B8)

S1 = |eˆ1iheˆ1| − |eˆ2iheˆ2|, (B9) 1 S2 = √ [|eˆ1iheˆ1| + |eˆ2iheˆ2| − 2|eˆ3iheˆ3|] , (B10) 3 S3 = |eˆ1iheˆ2| + |eˆ2iheˆ1|. (B11)

S4 = |eˆ1iheˆ3| + |eˆ3iheˆ1|. (B12)

S5 = |eˆ2iheˆ3| + |eˆ3iheˆ2|. (B13)

16 Here, we have employed the convention for the Gell-Mann matrices [13] λn and S1 = λ3,

S2 = λ8. Their matrix representations for eˆ1, eˆ2, and eˆ3 are   1 0 0   S   0 = 0 1 0 , (B14)   0 0 1   1 0 0   S   1 = 0 −1 0 , (B15)   0 0 0   1 0 0 1   S   2 = √ 0 1 0 , (B16) 3   0 0 −2   0 1 0   S   3 = 1 0 0 , (B17)   0 0 0   0 0 1   S   4 = 0 0 0 , (B18)   1 0 0   0 0 0   S   5 = 0 0 1 . (B19)   0 1 0

Note that only S0 is traceful, and that the remaining matrices are symmetric and traceless. The convention for the normalization is

Tr[S0] = 3, Tr[SiSj] = 2δij, (B20) for i, j = 1 through 5.

[1] See, for example, H. Goldstein, Classical Mechanics, 2nd ed. (Addison-Wesley, 1980); A. L. Fetter and J. D. Walecka, Theoretical Mechanics of Particles and Continua (Dover Publica-

17 tions, 2003); S. T. Thornton and J. B. Marion, Classical Dynamics of Particles and Systems, 4th ed. (Cengage Learning, 2003). [2] See, for example, R. Shankar, Principles of Quantum Mechanics, 2nd ed. (Plenum Press, New York, 1994); J. J. Sakurai and J. Napolitano, Modern Quantum Mechanics (Cambridge University Press, 2017). [3] W. R. Dixon, Am. J. Phys. 19, 536 (1951). [4] A. Rogers and P. Loly, Am. J. Phys. 72, 786 (2004). [5] A. Rogers and P. Loly, Eur. J. Phys. 26, 809 (2005). [6] J. Sivardiere, Eur. J. Phys. 4, 162 (1983). [7] F. Battaglia and T. F. George, Am. J. Phys. 81, 498 (2013). [8] A. R. Abdulghany, Am. J. Phys. 85, 791 (2017). [9] P. A. M. Dirac, Proc. Roy. Soc. A 113, 621 (1927). [10] P. A. M. Dirac, Math. Proc. Cambridge Philos. Soc., 35, 416 (1939). [11] P. A. M. Dirac, The Principles of Quantum Mechanics, 4th ed. (Clarendon Press, 1982). [12] J.-H. Ee, D.-W. Jung, U-R. Kim, and J. Lee, Eur. J. Phys. 38, 025801 (2017). [13] M. Gell-Mann, Phys. Rev. 125, 1067 (1962).

18