Tensor operator





In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.[1]


.mw-parser-output .toclimit-2 .toclevel-1 ul,.mw-parser-output .toclimit-3 .toclevel-2 ul,.mw-parser-output .toclimit-4 .toclevel-3 ul,.mw-parser-output .toclimit-5 .toclevel-4 ul,.mw-parser-output .toclimit-6 .toclevel-5 ul,.mw-parser-output .toclimit-7 .toclevel-6 ul{display:none}



Contents






  • 1 The general notion of scalar, vector, and tensor operators


  • 2 Rotations of quantum states


    • 2.1 Quantum rotation operator


    • 2.2 Angular momentum eigenkets




  • 3 Rotation of tensor operators


    • 3.1 Scalar operators


    • 3.2 Vector operators


    • 3.3 Spherical vector operators


    • 3.4 Tensor operators and their reducible and irreducible representations


    • 3.5 Spherical tensor operators




  • 4 Angular momentum and spherical harmonics


    • 4.1 Orbital angular momentum and spherical harmonics


    • 4.2 Spherical tensor operators and quantum spin




  • 5 Applications


    • 5.1 Dipole radiative transitions in a single-electron atom (alkali)


    • 5.2 Magnetic resonance


    • 5.3 Image processing and computer graphics




  • 6 See also


  • 7 References


    • 7.1 Notes


    • 7.2 Sources


    • 7.3 Further reading


      • 7.3.1 Spherical harmonics


      • 7.3.2 Angular momentum and spin


      • 7.3.3 Condensed matter physics


      • 7.3.4 Magnetic resonance


      • 7.3.5 Image processing






  • 8 External links





The general notion of scalar, vector, and tensor operators


In quantum mechanics, physical observables that are scalars, vectors, and tensors, must be represented by scalar, vector, and tensor operators, respectively. Whether something is a scalar, vector, or tensor depends on how it is viewed by two observers whose coordinate frames are related to each other by a rotation. Alternatively, one may ask how, for a single observer, a physical quantity transforms if the
state of the system is rotated. Consider, for example, a system consisting of a molecule of mass M{displaystyle M}M, traveling with a definite center of mass momentum, pz^{displaystyle p{mathbf {hat {z}} }}{displaystyle p{mathbf {hat {z}} }}, in the z{displaystyle z}z direction. If we rotate the system by 90∘{displaystyle 90^{circ }}{displaystyle 90^{circ }} about the y{displaystyle y}y axis, the momentum will change to px^{displaystyle p{mathbf {hat {x}} }}{displaystyle p{mathbf {hat {x}} }}, which is in the x{displaystyle x}x direction. The center-of-mass kinetic energy of the molecule will, however, be unchanged at p2/2M{displaystyle p^{2}/2M}{displaystyle p^{2}/2M}. The kinetic energy is a scalar and the momentum is a vector, and these two quantities must be represented by a scalar and a vector operator, respectively. By the latter in particular, we mean an operator whose expected values in the initial and the rotated states are pz^{displaystyle p{mathbf {hat {z}} }}{displaystyle p{mathbf {hat {z}} }} and px^{displaystyle p{mathbf {hat {x}} }}{displaystyle p{mathbf {hat {x}} }}. The kinetic energy on the other hand must be represented by a scalar operator, whose expected value must be the same in the initial and the rotated states.


In the same way, tensor quantities must be represented by tensor operators. An example of a tensor
quantity (of rank two) is the electrical quadrupole moment of the above molecule. Likewise, the octupole and hexadecapole moments would be tensors of rank three and four, respectively.


Other examples of scalar operators are the total energy operator (more commonly called the Hamiltonian), the potential energy, and the dipole-dipole interaction energy of two atoms. Examples of vector operators are the momentum, the position, the orbital angular momentum,
L{displaystyle {mathbf {L} }}{{mathbf  L}}, and the spin angular momentum, S{displaystyle {mathbf {S} }}{mathbf S}. (Fine print: Angular momentum is a vector as far as rotations are concerned, but unlike position or momentum it does not change sign under space inversion, and when one wishes to provide this information, it is said to be a pseudovector.)


Scalar, vector and tensor operators can also be formed by products of operators. For example, the scalar product L⋅S{displaystyle {mathbf {L} }cdot {mathbf {S} }}{displaystyle {mathbf {L} }cdot {mathbf {S} }} of the two vector operators, L{displaystyle {mathbf {L} }}{{mathbf  L}} and S{displaystyle {mathbf {S} }}{mathbf S}, is a scalar operator, which figures prominently in discussions of the spin-orbit interaction. Similarly, the quadrupole moment tensor of our example molecule has the nine components



Qij=∑α(3rα,irα,j−ij){displaystyle Q_{ij}=sum _{alpha }q_{alpha }(3r_{alpha ,i}r_{alpha ,j}-r_{alpha }^{2}delta _{ij})}{displaystyle Q_{ij}=sum _{alpha }q_{alpha }(3r_{alpha ,i}r_{alpha ,j}-r_{alpha }^{2}delta _{ij})}.

Here, the indices i{displaystyle i}i and j{displaystyle j}j can independently take on the values 1, 2, and 3 (or x{displaystyle x}x, y{displaystyle y}y, and z{displaystyle z}z) corresponding to the three Cartesian axes, the
index α{displaystyle alpha }alpha runs over all particles (electrons and nuclei) in the molecule,
{displaystyle q_{alpha }}{displaystyle q_{alpha }} is the charge on particle α{displaystyle alpha }alpha , and
,i{displaystyle r_{alpha ,i}}{displaystyle r_{alpha ,i}} is the i{displaystyle i}ith component of the position of this particle. Each term in the sum is a tensor operator. In particular, the nine products
,irα,j{displaystyle r_{alpha ,i}r_{alpha ,j}}{displaystyle r_{alpha ,i}r_{alpha ,j}} together form a second rank tensor, formed by taking the direct product of the vector operator {displaystyle {mathbf {r} }_{alpha }}{displaystyle {mathbf {r} }_{alpha }} with itself.



Rotations of quantum states



Quantum rotation operator


The rotation operator about the unit vector n (defining the axis of rotation) through angle θ is


U[R(θ,n^)]=exp⁡(−n^J){displaystyle U[R(theta ,{hat {mathbf {n} }})]=exp left(-{frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right)}{displaystyle U[R(theta ,{hat {mathbf {n} }})]=exp left(-{frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right)}

where J = (Jx, Jy, Jz) are the rotation generators (also the angular momentum matrices):


Jx=ℏ2(010101010)Jy=ℏ2(0i0−i0i0−i0)Jz=ℏ(−100000001){displaystyle J_{x}={frac {hbar }{sqrt {2}}}{begin{pmatrix}0&1&0\1&0&1\0&1&0end{pmatrix}},quad J_{y}={frac {hbar }{sqrt {2}}}{begin{pmatrix}0&i&0\-i&0&i\0&-i&0end{pmatrix}},quad J_{z}=hbar {begin{pmatrix}-1&0&0\0&0&0\0&0&1end{pmatrix}}}{displaystyle J_{x}={frac {hbar }{sqrt {2}}}{begin{pmatrix}0&1&0\1&0&1\0&1&0end{pmatrix}},quad J_{y}={frac {hbar }{sqrt {2}}}{begin{pmatrix}0&i&0\-i&0&i\0&-i&0end{pmatrix}},quad J_{z}=hbar {begin{pmatrix}-1&0&0\0&0&0\0&0&1end{pmatrix}}}

and let R^=R^,n^){displaystyle {widehat {R}}={widehat {R}}(theta ,{hat {mathbf {n} }})}widehat {R}=widehat {R}(theta ,{hat  {{mathbf  {n}}}}) be a rotation matrix. According to the Rodrigues' rotation formula, the rotation operator then amounts to


U[R(θ,n^)]=11−isin⁡θn^J−1−cos⁡θ2(n^J)2.{displaystyle U[R(theta ,{hat {mathbf {n} }})]=1!!1-{frac {isin theta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} -{frac {1-cos theta }{hbar ^{2}}}({hat {mathbf {n} }}cdot mathbf {J} )^{2}.}{displaystyle U[R(theta ,{hat {mathbf {n} }})]=1!!1-{frac {isin theta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} -{frac {1-cos theta }{hbar ^{2}}}({hat {mathbf {n} }}cdot mathbf {J} )^{2}.}

An operator Ω^{displaystyle {widehat {Omega }}}widehat {Omega } is invariant under a unitary transformation U if


Ω^=U†Ω^U;{displaystyle {widehat {Omega }}={U}^{dagger }{widehat {Omega }}U;}{displaystyle {widehat {Omega }}={U}^{dagger }{widehat {Omega }}U;}

in this case for the rotation U^(R){displaystyle {widehat {U}}(R)}widehat {U}(R),


Ω^=U(R)†Ω^U(R)=exp⁡(iθn^J)Ω^exp⁡(−n^J).{displaystyle {widehat {Omega }}={U(R)}^{dagger }{widehat {Omega }}U(R)=exp left({frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right){widehat {Omega }}exp left(-{frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right).}{displaystyle {widehat {Omega }}={U(R)}^{dagger }{widehat {Omega }}U(R)=exp left({frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right){widehat {Omega }}exp left(-{frac {itheta }{hbar }}{hat {mathbf {n} }}cdot mathbf {J} right).}


Angular momentum eigenkets


The orthonormal basis set for total angular momentum is |j,m⟩{displaystyle |j,mrangle }|j,mrangle , where j is the total angular momentum quantum number and m is the magnetic angular momentum quantum number, which takes values −j, −j + 1, ..., j − 1, j. A general state


=∑m|j,m⟩{displaystyle |psi rangle =sum _{m}|j,mrangle }|psi rangle =sum _{m}|j,mrangle

in the space rotates to a new state |j,m⟩{displaystyle |j,mrangle }|j,mrangle by:


¯=U(R)|ψ{displaystyle |{bar {psi }}rangle =U(R)|psi rangle }|{bar  {psi }}rangle =U(R)|psi rangle

Using the completeness condition:


I=∑m′|j,m′⟩j,m′|{displaystyle I=sum _{m'}|j,m'rangle langle j,m'|}I=sum _{{m'}}|j,m'rangle langle j,m'|

we have


¯=IU(R)|ψ=∑mm′|j,m′⟩j,m′|U(R)|j,m⟩{displaystyle |{bar {psi }}rangle =IU(R)|psi rangle =sum _{mm'}|j,m'rangle langle j,m'|U(R)|j,mrangle }|{bar  {psi }}rangle =IU(R)|psi rangle =sum _{{mm'}}|j,m'rangle langle j,m'|U(R)|j,mrangle

Introducing the Wigner D matrix elements:


D(R)m′m(j)=⟨j,m′|U(R)|j,m⟩{displaystyle {D(R)}_{m'm}^{(j)}=langle j,m'|U(R)|j,mrangle }{D(R)}_{{m'm}}^{{(j)}}=langle j,m'|U(R)|j,mrangle

gives the matrix multiplication:


¯=∑mm′Dm′m(j)|j,m′⟩¯=D(j)|ψ{displaystyle |{bar {psi }}rangle =sum _{mm'}D_{m'm}^{(j)}|j,m'rangle quad Rightarrow quad |{bar {psi }}rangle =D^{(j)}|psi rangle }{displaystyle |{bar {psi }}rangle =sum _{mm'}D_{m'm}^{(j)}|j,m'rangle quad Rightarrow quad |{bar {psi }}rangle =D^{(j)}|psi rangle }

For one basis ket:


|j,m¯=∑m′D(R)m′m(j)|j,m′⟩{displaystyle |{overline {j,m}}rangle =sum _{m'}{D(R)}_{m'm}^{(j)}|j,m'rangle }|overline {j,m}rangle =sum _{{m'}}{D(R)}_{{m'm}}^{{(j)}}|j,m'rangle

For the case of orbital angular momentum, the eigenstates |l,m⟩{displaystyle |l,mrangle }|l,mrangle of the orbital angular momentum operator L and solutions of Laplace's equation on a 3d sphere are spherical harmonics:


Yℓm(θ)=⟨θ|ℓ,m⟩=(2ℓ+1)4π(ℓm)!(ℓ+m)!Pℓm(cos⁡θ)eimϕ{displaystyle Y_{ell }^{m}(theta ,phi )=langle theta ,phi |ell ,mrangle ={sqrt {{(2ell +1) over 4pi }{(ell -m)! over (ell +m)!}}},P_{ell }^{m}(cos {theta }),e^{imphi }}{displaystyle Y_{ell }^{m}(theta ,phi )=langle theta ,phi |ell ,mrangle ={sqrt {{(2ell +1) over 4pi }{(ell -m)! over (ell +m)!}}},P_{ell }^{m}(cos {theta }),e^{imphi }}

where Pm is an associated Legendre polynomial, ℓ is the orbital angular momentum quantum number, and m is the orbital magnetic quantum number which takes the values −ℓ, −ℓ + 1, ... ℓ − 1, ℓ The formalism of spherical harmonics have wide applications in applied mathematics, and are closely related to the formalism of spherical tensors, as shown below.


Spherical harmonics are functions of the polar and azimuthal angles, ϕ and θ respectively, which can be conveniently collected into a unit vector n(θ, ϕ) pointing in the direction of those angles, in the Cartesian basis it is:


n^)=cos⁡ϕsin⁡θex+sin⁡ϕsin⁡θey+cos⁡θez{displaystyle {hat {mathbf {n} }}(theta ,phi )=cos phi sin theta mathbf {e} _{x}+sin phi sin theta mathbf {e} _{y}+cos theta mathbf {e} _{z}}{hat  {{mathbf  {n}}}}(theta ,phi )=cos phi sin theta {mathbf  {e}}_{x}+sin phi sin theta {mathbf  {e}}_{y}+cos theta {mathbf  {e}}_{z}

So a spherical harmonic can also be written Ylm=⟨n|lm⟩{displaystyle Y_{l}^{m}=langle mathbf {n} |lmrangle }Y_{{l}}^{{m}}=langle {mathbf  {n}}|lmrangle . Spherical harmonic states |m,l⟩{displaystyle |m,lrangle }|m,lrangle rotate according to the inverse rotation matrix U(R−1), while |l,m⟩{displaystyle |l,mrangle }|l,mrangle rotates by the initial rotation matrix U^(R){displaystyle {widehat {U}}(R)}widehat {U}(R).


|ℓ,m¯=∑m′Dm′m(ℓ)[U(R−1)]|ℓ,m′⟩,|n^¯=U(R)|n^{displaystyle |{overline {ell ,m}}rangle =sum _{m'}D_{m'm}^{(ell )}[U(R^{-1})]|ell ,m'rangle ,,quad |{overline {hat {mathbf {n} }}}rangle =U(R)|{hat {mathbf {n} }}rangle }{displaystyle |{overline {ell ,m}}rangle =sum _{m'}D_{m'm}^{(ell )}[U(R^{-1})]|ell ,m'rangle ,,quad |{overline {hat {mathbf {n} }}}rangle =U(R)|{hat {mathbf {n} }}rangle }


Rotation of tensor operators


We define the Rotation of an operator by requiring that the expectation value of the original operator A^{displaystyle {widehat {mathbf {A} }}}widehat {{mathbf  {A}}} with respect to the initial state be equal to the expectation value of the rotated operator with respect to the rotated state,


ψ′|A′^′⟩=⟨ψ|A^{displaystyle langle psi '|{widehat {A'}}|psi 'rangle =langle psi |{widehat {A}}|psi rangle }langle psi '|widehat {A'}|psi 'rangle =langle psi |widehat {A}|psi rangle

Now as,



{displaystyle |psi rangle } | psi rangle ′⟩=U(R)|ψ,⟨ψ|{displaystyle |psi 'rangle =U(R)|psi rangle ,,quad langle psi |}|psi 'rangle =U(R)|psi rangle ,,quad langle psi |ψ′|=⟨ψ|U†(R){displaystyle langle psi '|=langle psi |U^{dagger }(R)}langle psi '|=langle psi |U^{dagger }(R)

we have,


ψ|U†(R)A^′U(R)|ψ=⟨ψ|A^{displaystyle langle psi |U^{dagger }(R){widehat {A}}'U(R)|psi rangle =langle psi |{widehat {A}}|psi rangle }langle psi |U^{dagger }(R)widehat {A}'U(R)|psi rangle =langle psi |widehat {A}|psi rangle

since, {displaystyle |psi rangle } | psi rangle is arbitrary,


U†(R)A^′U(R)=A^{displaystyle U^{dagger }(R){widehat {A}}'U(R)={widehat {A}}}U^{dagger }(R)widehat {A}'U(R)=widehat {A}


Scalar operators


A scalar operator is invariant under rotations:[2]


U(R)†S^U(R)=S^{displaystyle U(R)^{dagger }{widehat {S}}U(R)={widehat {S}}}U(R)^{dagger }widehat {S}U(R)=widehat {S}

and we have a simple result, that the scalar operator commutes with the rotation generators:


[S^,J^]=0{displaystyle left[{widehat {S}},{widehat {mathbf {J} }}right]=0}left[widehat {S},widehat {{mathbf  {J}}}right]=0

Examples of a scalar operators include


  • the energy operator:

E^ψ=iℏ{displaystyle {widehat {E}}psi =ihbar {frac {partial }{partial t}}psi }widehat {E}psi =ihbar {frac  {partial }{partial t}}psi


  • potential energy V

V^(r,t)ψ=V(r,t)ψ{displaystyle {widehat {V}}(mathbf {r} ,t)psi =V(mathbf {r} ,t)psi }widehat {V}({mathbf  {r}},t)psi =V({mathbf  {r}},t)psi


  • kinetic energy T:

T^(r,t)ψ=−22m∇2(r,t){displaystyle {widehat {T}}(mathbf {r} ,t)psi =-{frac {hbar ^{2}}{2m}}nabla ^{2}(mathbf {r} ,t)}widehat {T}({mathbf  {r}},t)psi =-{frac  {hbar ^{2}}{2m}}nabla ^{2}({mathbf  {r}},t)

  • as well as the operator coupling term in spin-orbit coupling:

L^S^=L^xS^x+L^yS^y+L^zS^z.{displaystyle {widehat {mathbf {L} }}cdot {widehat {mathbf {S} }}={widehat {L}}_{x}{widehat {S}}_{x}+{widehat {L}}_{y}{widehat {S}}_{y}+{widehat {L}}_{z}{widehat {S}}_{z},.}widehat {{mathbf  {L}}}cdot widehat {{mathbf  {S}}}=widehat {L}_{x}widehat {S}_{x}+widehat {L}_{y}widehat {S}_{y}+widehat {L}_{z}widehat {S}_{z},.


Vector operators


Vector operators (as well as pseudovector operators) are a set of 3 operators that can be rotated according to:[2]


U(R)†V^iU(R)=∑jRijV^j{displaystyle {U(R)}^{dagger }{widehat {V}}_{i}U(R)=sum _{j}R_{ij}{widehat {V}}_{j}}{U(R)}^{dagger }widehat {V}_{i}U(R)=sum _{j}R_{{ij}}widehat {V}_{j}

from this and the infinitesimal rotation operator and its Hermitian conjugate, and ignoring second order term in θ)2{displaystyle (delta theta )^{2}}{displaystyle (delta theta )^{2}},one can derive the commutation relation with the rotation generator:


[V^a,J^b]≈iℏεabcV^c{displaystyle left[{widehat {V}}_{a},{widehat {J}}_{b}right]approx ihbar varepsilon _{abc}{widehat {V}}_{c}}{displaystyle left[{widehat {V}}_{a},{widehat {J}}_{b}right]approx ihbar varepsilon _{abc}{widehat {V}}_{c}}

where εijk is the Levi-Civita symbol, which all vector operators must satisfy, by construction. As the symbol εijk is a pseudotensor, pseudovector operators are invariant up to a sign: +1 for proper rotations and −1 for improper rotations.


Vector operators include


  • the position operator:

r^ψ=rψ{displaystyle {widehat {mathbf {r} }}psi =mathbf {r} psi }widehat {{mathbf  {r}}}psi ={mathbf  {r}}psi

  • the momentum operator:

p^ψ=−iℏψ{displaystyle {widehat {mathbf {p} }}psi =-ihbar nabla psi }widehat {{mathbf  {p}}}psi =-ihbar nabla psi

and peusodovector operators include


  • the orbital angular momentum operator:

L^ψ=−iℏψ{displaystyle {widehat {mathbf {L} }}psi =-ihbar mathbf {r} times nabla psi }widehat {{mathbf  {L}}}psi =-ihbar {mathbf  {r}}times nabla psi

  • as well the spin operator S, and hence the total angular momentum

J^=L^+S^.{displaystyle {widehat {mathbf {J} }}={widehat {mathbf {L} }}+{widehat {mathbf {S} }},.}widehat {{mathbf  {J}}}=widehat {{mathbf  {L}}}+widehat {{mathbf  {S}}},.

In Dirac notation:


ψ¯|V^a|ψ¯=⟨ψ|U(R)†V^aU(R)|ψ=∑bRab⟨ψ|V^b|ψ{displaystyle langle {bar {psi }}|{widehat {V}}_{a}|{bar {psi }}rangle =langle psi |{U(R)}^{dagger }{widehat {V}}_{a}U(R)|psi rangle =sum _{b}R_{ab}langle psi |{widehat {V}}_{b}|psi rangle }langle {bar  {psi }}|widehat {V}_{a}|{bar  {psi }}rangle =langle psi |{U(R)}^{dagger }widehat {V}_{a}U(R)|psi rangle =sum _{b}R_{{ab}}langle psi |widehat {V}_{b}|psi rangle

and since | Ψ > is any quantum state, the same result follows:


U(R)†V^aU(R)=∑bRabV^b{displaystyle {U(R)}^{dagger }{widehat {V}}_{a}U(R)=sum _{b}R_{ab}{widehat {V}}_{b}}{U(R)}^{dagger }widehat {V}_{a}U(R)=sum _{b}R_{{ab}}widehat {V}_{b}

Note that here, the term "vector" is used two different ways: kets such as |ψ are elements of abstract Hilbert spaces, while the vector operator is defined as a quantity whose components transform in a certain way under rotations.



Spherical vector operators


A vector operator in the spherical basis is V = (V+1, V0, V−1) where the components are:[2]


V+1=−12(Vx+iVy)V−1=12(Vx−iVy),V0=Vz,{displaystyle V_{+1}=-{frac {1}{sqrt {2}}}(V_{x}+iV_{y}),quad V_{-1}={frac {1}{sqrt {2}}}(V_{x}-iV_{y}),,quad V_{0}=V_{z},,}V_{{+1}}=-{frac  {1}{{sqrt  {2}}}}(V_{x}+iV_{y}),quad V_{{-1}}={frac  {1}{{sqrt  {2}}}}(V_{x}-iV_{y}),,quad V_{0}=V_{z},,

and the commutators with the rotation generators are:



[Jz,Vq]=qVq{displaystyle left[J_{z},V_{q}right]=qV_{q}}left[J_{z},V_{q}right]=qV_{q}

[J±,V0]=2V±{displaystyle left[J_{pm },V_{0}right]={sqrt {2}}V_{pm }}left[J_{pm },V_{0}right]={sqrt  {2}}V_{pm }

[J±,V∓]=2V0{displaystyle left[J_{pm },V_{mp }right]={sqrt {2}}V_{0}}left[J_{pm },V_{mp }right]={sqrt  {2}}V_{0}

[J±,V±]=0{displaystyle left[J_{pm },V_{pm }right]=0}left[J_{pm },V_{pm }right]=0


where q is a placeholder for the spherical basis labels (+1, 0, −1), and:


=Jx±iJy,{displaystyle J_{pm }=J_{x}pm iJ_{y},,}J_{pm }=J_{x}pm iJ_{y},,

(some authors may place a factor of 1/2 on the left hand side of the equation) and raise (J+) or lower (J) the total magnetic quantum number m by one unit. In the spherical basis the generators are:


1=∓12J±,J0=Jz{displaystyle J_{pm 1}=mp {frac {1}{sqrt {2}}}J_{pm },,quad J_{0}=J_{z}}J_{{pm 1}}=mp {frac  {1}{{sqrt  {2}}}}J_{pm },,quad J_{0}=J_{z}

The rotation transformation in the spherical basis (originally written in the Cartesian basis) is then:


U(R)†V^qU(R)=∑q′D(R)qq′(1)V^q′{displaystyle {U(R)}^{dagger }{widehat {V}}_{q}U(R)=sum _{q'}{D(R)}_{qq'}^{(1)}{widehat {V}}_{q'}}{U(R)}^{dagger }widehat {V}_{q}U(R)=sum _{{q'}}{D(R)}_{{qq'}}^{{(1)}}widehat {V}_{{q'}}

One can generalize the vector operator concept easily to tensorial operators, shown next.



Tensor operators and their reducible and irreducible representations


A tensor operator can be rotated according to:[2]


U(R)†T^pqr⋯U(R)=RpiRqjRrk⋯T^ijk⋯{displaystyle U(R)^{dagger }{widehat {T}}_{pqrcdots }U(R)=R_{pi}R_{qj}R_{rk}cdots {widehat {T}}_{ijkcdots }}U(R)^{dagger }widehat {T}_{{pqrcdots }}U(R)=R_{{pi}}R_{{qj}}R_{{rk}}cdots widehat {T}_{{ijkcdots }}

Consider a dyadic tensor with components Tij = aibj, this rotates infinitesimally according to:


U(R)†T^pqU(R)=RpiRqjT^ij=Rpia^iRqjb^j{displaystyle U(R)^{dagger }{widehat {T}}_{pq}U(R)=R_{pi}R_{qj}{widehat {T}}_{ij}=R_{pi}{widehat {a}}_{i}R_{qj}{widehat {b}}_{j}}U(R)^{dagger }widehat {T}_{{pq}}U(R)=R_{{pi}}R_{{qj}}widehat {T}_{{ij}}=R_{{pi}}widehat {a}_{i}R_{{qj}}widehat {b}_{j}

Cartesian dyadic tensors of the form


T^=eia^i⊗ejb^j=ei⊗eja^ib^j{displaystyle {hat {mathbf {T} }}=mathbf {e} _{i}{widehat {a}}_{i}otimes mathbf {e} _{j}{widehat {b}}_{j}=mathbf {e} _{i}otimes mathbf {e} _{j}{widehat {a}}_{i}{widehat {b}}_{j}}{hat  {{mathbf  {T}}}}={mathbf  {e}}_{i}widehat {a}_{i}otimes {mathbf  {e}}_{j}widehat {b}_{j}={mathbf  {e}}_{i}otimes {mathbf  {e}}_{j}widehat {a}_{i}widehat {b}_{j}

where a and b are two vector operators:


a^=eia^i,b^=ejb^j{displaystyle {hat {mathbf {a} }}=mathbf {e} _{i}{widehat {a}}_{i},,quad {hat {mathbf {b} }}=mathbf {e} _{j}{widehat {b}}_{j}}{hat  {{mathbf  {a}}}}={mathbf  {e}}_{i}widehat {a}_{i},,quad {hat  {{mathbf  {b}}}}={mathbf  {e}}_{j}widehat {b}_{j}

are reducible, which means they can be re-expressed in terms of a and b as a rank 0 tensor (scalar), plus a rank 1 tensor (an antisymmetric tensor), plus a rank 2 tensor (a symmetric tensor with zero trace):


T=T(0)+T(1)+T(2){displaystyle mathbf {T} =mathbf {T} ^{(0)}+mathbf {T} ^{(1)}+mathbf {T} ^{(2)}}{displaystyle mathbf {T} =mathbf {T} ^{(0)}+mathbf {T} ^{(1)}+mathbf {T} ^{(2)}}

where the first term


T^ij(0)=a^kb^k3δij{displaystyle {widehat {T}}_{ij}^{(0)}={frac {{widehat {a}}_{k}{widehat {b}}_{k}}{3}}delta _{ij}}{displaystyle {widehat {T}}_{ij}^{(0)}={frac {{widehat {a}}_{k}{widehat {b}}_{k}}{3}}delta _{ij}}

includes just one component, a scalar equivalently written (a·b)/3, the second


T^ij(1)=12[a^ib^j−a^jb^i]=a^[ib^j]{displaystyle {widehat {T}}_{ij}^{(1)}={frac {1}{2}}[{widehat {a}}_{i}{widehat {b}}_{j}-{widehat {a}}_{j}{widehat {b}}_{i}]={widehat {a}}_{[i}{widehat {b}}_{j]}}{displaystyle {widehat {T}}_{ij}^{(1)}={frac {1}{2}}[{widehat {a}}_{i}{widehat {b}}_{j}-{widehat {a}}_{j}{widehat {b}}_{i}]={widehat {a}}_{[i}{widehat {b}}_{j]}}

includes three independent components, equivalently the components of (a×b)/2, and the third


T^ij(2)=12(a^ib^j+a^jb^i)−a^kb^k3δij=a^(ib^j)−Tij(0){displaystyle {widehat {T}}_{ij}^{(2)}={frac {1}{2}}({widehat {a}}_{i}{widehat {b}}_{j}+{widehat {a}}_{j}{widehat {b}}_{i})-{frac {{widehat {a}}_{k}{widehat {b}}_{k}}{3}}delta _{ij}={widehat {a}}_{(i}{widehat {b}}_{j)}-T_{ij}^{(0)}}{displaystyle {widehat {T}}_{ij}^{(2)}={frac {1}{2}}({widehat {a}}_{i}{widehat {b}}_{j}+{widehat {a}}_{j}{widehat {b}}_{i})-{frac {{widehat {a}}_{k}{widehat {b}}_{k}}{3}}delta _{ij}={widehat {a}}_{(i}{widehat {b}}_{j)}-T_{ij}^{(0)}}

includes five independent components. Throughout, δij is the Kronecker delta, the components of the identity matrix. The number in the superscripted brackets denotes the tensor rank. These three terms are irreducible, which means they cannot be decomposed further and still be tensors satisfying the defining transformation laws under which they must be invariant. These also correspond to the number of spherical harmonic functions 2ℓ + 1 for ℓ = 0, 1, 2, the same as the ranks for each tensor. Each of the irreducible representations T(1), T(2) ... transform like angular momentum eigenstates according to the number of independent components.


Example of a Tensor operator,


  • The Quadrupole moment operator,

Qij=∑α(3rαirαj−ij){displaystyle Q_{ij}=sum _{alpha }q_{alpha }(3r_{alpha i}r_{alpha j}-r_{alpha }^{2}delta _{ij})}Q_{{ij}}=sum _{{alpha }}q_{{alpha }}(3r_{{alpha i}}r_{{alpha j}}-r_{{alpha }}^{{2}}delta _{{ij}})

  • Two Tensor operators can be multiplied to give another Tensor operator.

Tij=ViWj{displaystyle T_{ij}=V_{i}W_{j}}T_{{ij}}=V_{{i}}W_{{j}}

in general,


Ti1i2⋯j1j2⋯=Vi1i2⋯Wj1j2⋯{displaystyle T_{i_{1}i_{2}cdots j_{1}j_{2}cdots }=V_{i_{1}i_{2}cdots }W_{j_{1}j_{2}cdots }}T_{{i_{{1}}i_{{2}}cdots j_{{1}}j_{{2}}cdots }}=V_{{i_{{1}}i_{{2}}cdots }}W_{{j_{{1}}j_{{2}}cdots }}

Note: This is just an example, in general, a tensor operator cannot be written as the product of two Tensor operators as given in the above example.



Spherical tensor operators


Continuing the previous example of the second order dyadic tensor T = ab, casting each of a and b into the spherical basis and substituting into T gives the spherical tensor operators of the second order, which are:



T^±2(2)=a^±1b^±1{displaystyle {widehat {T}}_{pm 2}^{(2)}={widehat {a}}_{pm 1}{widehat {b}}_{pm 1}}widehat {T}_{{pm 2}}^{{(2)}}=widehat {a}_{{pm 1}}widehat {b}_{{pm 1}}

T^±1(2)=12(a^±1b^0+a^0b^±1){displaystyle {widehat {T}}_{pm 1}^{(2)}={frac {1}{sqrt {2}}}left({widehat {a}}_{pm 1}{widehat {b}}_{0}+{widehat {a}}_{0}{widehat {b}}_{pm 1}right)}widehat {T}_{{pm 1}}^{{(2)}}={frac  {1}{{sqrt  {2}}}}left(widehat {a}_{{pm 1}}widehat {b}_{0}+widehat {a}_{0}widehat {b}_{{pm 1}}right)

T^0(2)=16(a^+1b^1+a^1b^+1+2a^0b^0){displaystyle {widehat {T}}_{0}^{(2)}={frac {1}{sqrt {6}}}left({widehat {a}}_{+1}{widehat {b}}_{-1}+{widehat {a}}_{-1}{widehat {b}}_{+1}+2{widehat {a}}_{0}{widehat {b}}_{0}right)}widehat {T}_{{0}}^{{(2)}}={frac  {1}{{sqrt  {6}}}}left(widehat {a}_{{+1}}widehat {b}_{{-1}}+widehat {a}_{{-1}}widehat {b}_{{+1}}+2widehat {a}_{0}widehat {b}_{0}right)


Using the infinitesimal rotation operator and its Hermitian conjugate, one can derive the commutation relation in the spherical basis:


[Ja,T^q(2)]=∑q′D(Ja)qq′(2)T^q′(2)=∑q′⟨j=2,m=q|Ja|j=2,m=q′⟩T^q′(2){displaystyle left[J_{a},{widehat {T}}_{q}^{(2)}right]=sum _{q'}{D(J_{a})}_{qq'}^{(2)}{widehat {T}}_{q'}^{(2)}=sum _{q'}langle j{=}2,m{=}q|J_{a}|j{=}2,m{=}q'rangle {widehat {T}}_{q'}^{(2)}}{displaystyle left[J_{a},{widehat {T}}_{q}^{(2)}right]=sum _{q'}{D(J_{a})}_{qq'}^{(2)}{widehat {T}}_{q'}^{(2)}=sum _{q'}langle j{=}2,m{=}q|J_{a}|j{=}2,m{=}q'rangle {widehat {T}}_{q'}^{(2)}}

and the finite rotation transformation in the spherical basis is:


U(R)†T^q(2)U(R)=∑q′D(R)qq′(2)T^q′(2){displaystyle {U(R)}^{dagger }{widehat {T}}_{q}^{(2)}U(R)=sum _{q'}{D(R)}_{qq'}^{(2)}{widehat {T}}_{q'}^{(2)}}{U(R)}^{dagger }widehat {T}_{q}^{{(2)}}U(R)=sum _{{q'}}{D(R)}_{{qq'}}^{{(2)}}widehat {T}_{{q'}}^{{(2)}}

In general, tensor operators can be constructed from two perspectives.[3]


One way is to specify how spherical tensors transform under a physical rotation - a group theoretical definition. A rotated angular momentum eigenstate can be decomposed into a linear combination of the initial eigenstates: the coefficients in the linear combination consist of Wigner rotation matrix entries. Spherical tensor operators are sometimes defined as the set of operators that transform just like the eigenkets under a rotation.


A spherical tensor Tq(k) of rank k is defined to rotate into Tq(k) according to:


U(R)†T^q(k)U(R)=∑q′D(R)qq′(k)T^q′(k){displaystyle {U(R)}^{dagger }{widehat {T}}_{q}^{(k)}U(R)=sum _{q'}D(R)_{qq'}^{(k)}{widehat {T}}_{q'}^{(k)}}{displaystyle {U(R)}^{dagger }{widehat {T}}_{q}^{(k)}U(R)=sum _{q'}D(R)_{qq'}^{(k)}{widehat {T}}_{q'}^{(k)}}

where q = k, k − 1, ..., −k + 1, −k. For spherical tensors, k and q are analogous labels to ℓ and m respectively, for spherical harmonics. Some authors write Tkq instead of Tq(k), with or without the parentheses enclosing the rank number k.


Another related procedure requires that the spherical tensors satisfy certain commutation relations with respect to the rotation generators Jx, Jy, Jz - an algebraic definition.


The commutation relations of the angular momentum components with the tensor operators are:



[J±,T^q(k)]=ℏ(k∓q)(k±q+1)T^1(k){displaystyle left[J_{pm },{widehat {T}}_{q}^{(k)}right]=hbar {sqrt {(kmp q)(kpm q+1)}}{widehat {T}}_{qpm 1}^{(k)}}{displaystyle left[J_{pm },{widehat {T}}_{q}^{(k)}right]=hbar {sqrt {(kmp q)(kpm q+1)}}{widehat {T}}_{qpm 1}^{(k)}}

[Jz,T^q(k)]=ℏqT^q(k){displaystyle left[J_{z},{widehat {T}}_{q}^{(k)}right]=hbar q{widehat {T}}_{q}^{(k)}}{displaystyle left[J_{z},{widehat {T}}_{q}^{(k)}right]=hbar q{widehat {T}}_{q}^{(k)}}


For any 3d vector, not just a unit vector, and not just the position vector:


a=axex+ayey+azez{displaystyle mathbf {a} =a_{x}mathbf {e} _{x}+a_{y}mathbf {e} _{y}+a_{z}mathbf {e} _{z}}{mathbf  {a}}=a_{x}{mathbf  {e}}_{x}+a_{y}{mathbf  {e}}_{y}+a_{z}{mathbf  {e}}_{z}

a spherical tensor is a spherical harmonic as a function of this vector a, and in Dirac notation:


Tq(k)=Yℓ=km=q(a)=⟨a|k,q⟩{displaystyle T_{q}^{(k)}=Y_{ell =k}^{m=q}(mathbf {a} )=langle mathbf {a} |k,qrangle }T_{q}^{{(k)}}=Y_{{ell =k}}^{{m=q}}({mathbf  {a}})=langle {mathbf  {a}}|k,qrangle

(the super and subscripts switch places for the corresponding labels k and mq which spherical tensors and spherical harmonics use).


Spherical harmonic states and spherical tensors can also be constructed out of the Clebsch–Gordan coefficients. Irreducible spherical tensors can build higher rank spherical tensors; if Aq1(k1) and Bq2(k2) are two spherical tensors of ranks k1 and k2 respectively, then:


Tq(k)=∑q1q2⟨k1,k2,q1,q2|k,q,q1,q2⟩Aq1(k1)Bq2(k2){displaystyle T_{q}^{(k)}=sum _{q_{1}q_{2}}langle k_{1},k_{2},q_{1},q_{2}|k,q,q_{1},q_{2}rangle A_{q_{1}}^{(k_{1})}B_{q_{2}}^{(k_{2})}}{displaystyle T_{q}^{(k)}=sum _{q_{1}q_{2}}langle k_{1},k_{2},q_{1},q_{2}|k,q,q_{1},q_{2}rangle A_{q_{1}}^{(k_{1})}B_{q_{2}}^{(k_{2})}}

is a spherical tensor of rank k.


The Hermitian adjoint of a spherical tensor may be defined as


(T†)q(k)=(−1)k−q(T−q(k))†.{displaystyle (T^{dagger })_{q}^{(k)}=(-1)^{k-q}(T_{-q}^{(k)})^{dagger }.}{displaystyle (T^{dagger })_{q}^{(k)}=(-1)^{k-q}(T_{-q}^{(k)})^{dagger }.}

There is some arbitrariness in the choice of the phase factor: any factor containing (−1)±q will satisfy the commutation relations.[4] The above choice of phase has the advantages of being real and that the tensor product two commuting Hermitian operators is still Hermitian.[5] Some authors define it with a different sign on q, without the k, or use only the floor of k[6].



Angular momentum and spherical harmonics



Orbital angular momentum and spherical harmonics


Orbital angular momentum operators have the ladder operators:


=Lx±iLy{displaystyle L_{pm }=L_{x}pm iL_{y}}L_{pm }=L_{x}pm iL_{y}

which raise or lower the orbital magnetic quantum number m by one unit. This has almost exactly the same form as the spherical basis, aside from constant multiplicative factors.



Spherical tensor operators and quantum spin



Spherical tensors can also be formed from algebraic combinations of the spin operators Sx, Sy, Sz, as matrices, for a spin system with total quantum number j = ℓ + s (and ℓ = 0). Spin operators have the ladder operators:


=Sx±iSy{displaystyle S_{pm }=S_{x}pm iS_{y}}S_{pm }=S_{x}pm iS_{y}

which raise or lower the spin magnetic quantum number ms by one unit.



Applications


Spherical bases have broad applications in pure and applied mathematics and physical sciences where spherical geometries occur.



Dipole radiative transitions in a single-electron atom (alkali)


The transition amplitude is proportional to matrix elements of the dipole operator between the initial and final states. We use an electrostatic, spinless model for the atom and we consider the transition from the initial energy level Enℓ to final level En′ℓ′. These levels are degenerate, since the energy does not depend on the magnetic quantum number m or m′. The wave functions have the form,


ψnlm(r,θ)=Rnl(r)Ylm(θ){displaystyle psi _{nlm}(r,theta ,phi )=R_{nl}(r)Y_{lm}(theta ,phi )}psi _{{nlm}}(r,theta ,phi )=R_{{nl}}(r)Y_{{lm}}(theta ,phi )

The dipole operator is proportional to the position operator of the electron, so we must evaluate matrix elements of the form,


n′l′m′|r|nlm⟩{displaystyle langle n'l'm'|mathbf {r} |nlmrangle }langle n'l'm'|{mathbf  {r}}|nlmrangle

where, the initial state is on the right and the final one on the left. The position operator r has three components, and the initial and final levels consist of 2ℓ + 1 and 2ℓ′ + 1 degenerate states, respectively. Therefore if we wish to evaluate the intensity of a spectral line as it would be observed, we really have to evaluate 3(2ℓ′+ 1)(2ℓ+ 1) matrix elements, for example, 3×3×5 = 45 in a 3d → 2p transition. This is actually an exaggeration, as we shall see, because many of the matrix elements vanish, but there are still many non-vanishing matrix elements to be calculated.


A great simplification can be achieved by expressing the components of r, not with respect to the Cartesian basis, but with respect to the spherical basis. First we define,


rq=e^q⋅r{displaystyle r_{q}={hat {mathbf {e} }}_{q}cdot mathbf {r} }r_{{q}}={hat  {{mathbf  {e}}}}_{{q}}cdot {mathbf  {r}}

Next, by inspecting a table of the Yℓm′s, we find that for ℓ = 1 we have,



rY11(θ)=−r38πsin⁡)eiϕ=34π(−x+iy2){displaystyle rY_{11}(theta ,phi )=-r{sqrt {frac {3}{8pi }}}sin(theta )e^{iphi }={sqrt {frac {3}{4pi }}}left(-{frac {x+iy}{sqrt {2}}}right)}rY_{{11}}(theta ,phi )=-r{sqrt  {{frac  {3}{8pi }}}}sin(theta )e^{{iphi }}={sqrt  {{frac  {3}{4pi }}}}left(-{frac  {x+iy}{{sqrt  {2}}}}right)

rY10(θ)=r34πcos⁡)=34πz{displaystyle rY_{10}(theta ,phi )=r{sqrt {frac {3}{4pi }}}cos(theta )={sqrt {frac {3}{4pi }}}z}{displaystyle rY_{10}(theta ,phi )=r{sqrt {frac {3}{4pi }}}cos(theta )={sqrt {frac {3}{4pi }}}z}

rY1−1(θ)=r38πsin⁡)e−=34π(x−iy2){displaystyle rY_{1-1}(theta ,phi )=r{sqrt {frac {3}{8pi }}}sin(theta )e^{-iphi }={sqrt {frac {3}{4pi }}}left({frac {x-iy}{sqrt {2}}}right)}rY_{{1-1}}(theta ,phi )=r{sqrt  {{frac  {3}{8pi }}}}sin(theta )e^{{-iphi }}={sqrt  {{frac  {3}{4pi }}}}left({frac  {x-iy}{{sqrt  {2}}}}right)


where, we have multiplied each Y1m by the radius r. On the right hand side we see the spherical components rq of the position vector r. The results can be summarized by,


rY1q(θ)=34πrq{displaystyle rY_{1q}(theta ,phi )={sqrt {frac {3}{4pi }}}r_{q}}rY_{{1q}}(theta ,phi )={sqrt  {{frac  {3}{4pi }}}}r_{{q}}

for q = 1, 0, −1, where q appears explicitly as a magnetic quantum number. This equation reveals a relationship between vector operators and the angular momentum value ℓ = 1, something we will have more to say about presently. Now the matrix elements become a product of a radial integral times an angular integral,


n′l′m′|r|nlm⟩=(∫0∞r2drRn′l′∗(r)rRnl(r))(4π3∫Yl′m′∗)Y1q(θ)Ylm(θ)){displaystyle langle n'l'm'|mathbf {r} |nlmrangle =left(int _{0}^{infty }r^{2}drR_{n'l'}^{*}(r)rR_{nl}(r)right)left({sqrt {frac {4pi }{3}}}int dOmega Y_{l'm'}^{*}(theta ,phi )Y_{1q}(theta ,phi )Y_{lm}(theta ,phi )right)}langle n'l'm'|{mathbf  {r}}|nlmrangle =left(int _{0}^{{infty }}r^{2}drR_{{n'l'}}^{*}(r)rR_{{nl}}(r)right)left({sqrt  {{frac  {4pi }{3}}}}int dOmega Y_{{l'm'}}^{*}(theta ,phi )Y_{{1q}}(theta ,phi )Y_{{lm}}(theta ,phi )right)

We see that all the dependence on the three magnetic quantum numbers (m′,q,m) is contained in the angular part of the integral. Moreover, the angular integral can be evaluated by the three-Yℓm formula, whereupon it becomes proportional to the Clebsch-Gordan coefficient,


l′m′|l1mq⟩{displaystyle langle l'm'|l1mqrangle }langle l'm'|l1mqrangle

The radial integral is independent of the three magnetic quantum numbers (m′, q, m), and the trick we have just used does not help us to evaluate it. But it is only one integral, and after it has been done, all the other integrals can be evaluated just by computing or looking up Clebsch-Gordan
coefficients.


The selection rule m′ = q + m in the Clebsch-Gordan coefficient means that many of the integrals vanish, so we have exaggerated the total number of integrals that need to be done. But had we worked with the Cartesian components ri of r, this selection rule might not have been obvious. In any case, even with the selection rule, there may still be many nonzero integrals to be done (nine, in the case 3d → 2p).
The example we have just given of simplifying the calculation of matrix elements for a dipole transition is really an application of the Wigner-Eckart theorem, which we take up later in these notes.



Magnetic resonance


The spherical tensor formalism provides a common platform for treating coherence and relaxation in nuclear magnetic resonance. In NMR and EPR, spherical tensor operators are employed to express the quantum dynamics of particle spin, by means of an equation of motion for the density matrix entries, or to formulate dynamics in terms of an equation of motion in Liouville space. The Liouville space equation of motion governs the observable averages of spin variables. When relaxation is formulated using a spherical tensor basis in Liouville space, insight is gained because the relaxation matrix exhibits the cross-relaxation of spin observables directly.[3]



Image processing and computer graphics



See also



  • Wigner–Eckart theorem

  • Structure tensor

  • Clebsch-Gordan coefficient for SU(3)



References



Notes





  1. ^ Jeevanjee, Nadir (2015). An Introduction to Tensors and Group Theory for Physicists (2nd ed.). Birkhauser. ISBN 978-0-8176-4714-8..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  2. ^ abcd E. Abers (2004). "5". Quantum Mechanics. Addison Wesley. ISBN 978-0-13-146100-0.


  3. ^ ab R.D. Nielsen; B.H. Robinson (2006). "The Spherical Tensor Formalism Applied to Relaxation in Magnetic Resonance" (PDF). pp. 270–271.


  4. ^ McCarthy, Ian E.; Weigold, Erich (2005). Electron-Atom Collisions (Volume 5 of Cambridge Monographs on Atomic, Molecular and Chemical Physics). Cambridge University Press. p. 68. ISBN 9780521019682.


  5. ^ Edmonds, A. R. (1957). Angular Momentum in Quantum Mechanics. Princeton University Press. p. 78. ISBN 9780691025896.


  6. ^ Degl'Innocenti, M. Landi; Landolfi, M. (2006). Polarization in Spectral Lines. Springer Science & Business Media. p. 65. ISBN 9781402024153.




Sources


.mw-parser-output .refbegin{font-size:90%;margin-bottom:0.5em}.mw-parser-output .refbegin-hanging-indents>ul{list-style-type:none;margin-left:0}.mw-parser-output .refbegin-hanging-indents>ul>li,.mw-parser-output .refbegin-hanging-indents>dl>dd{margin-left:0;padding-left:3.2em;text-indent:-3.2em;list-style:none}.mw-parser-output .refbegin-100{font-size:100%}



  • P. T. Callaghan (2011). Translational Dynamics and Magnetic Resonance:Principles of Pulsed Gradient Spin Echo NMR. Oxford University Press. ISBN 978-0-191-621-048.


  • V. V. Balashov; A. N. Grum-Grzhimailo; N.M. Kabachnik (2000). Polarization and Correlation Phenomena in Atomic Collisions: A Practical Theory Course. Springer. ISBN 9780306462665.


  • J. A. Tuszynski (1990). Spherical Tensor Operators: Tables of Matrix Elements and Symmetries. World Scientific. ISBN 978-981-0202-835.


  • L. Castellani; J. Wess (1996). Quantum Groups and Their Applications in Physics: Varenna on Lake Como, Villa Monastero, 28 June-8 July 1994. Società Italiana di Fisica, IOS. ISBN 978-905-199-24-72.


  • Introduction to the Graphical Theory of Angular Momentum. Springer. 2009. ISBN 978-364-203-11-99.


  • A. R. Edmonds (1996). Angular Momentum in Quantum Mechanics (2nd ed.). Princeton University Press. ISBN 978-0-691-025-896.


  • L.J. Mueller (2011). "Tensors and rotations in NMR". Concepts in Magnetic Resonance Part A. 38A (5): 221–235. doi:10.1002/cmr.a.20224.


  • M.S. Anwar (2004). "Spherical Tensor Operators in NMR" (PDF).


  • P. Callaghan (1993). Principles of nuclear magnetic resonance microscopy. Oxford University Press. pp. 56–57. ISBN 978-0-198-539-971.




Further reading



Spherical harmonics




  • G.W.F. Drake (2006). Springer Handbook of Atomic, Molecular, and Optical Physics (2nd ed.). Springer. p. 57. ISBN 978-0-3872-6308-3.


  • F.A. Dahlen; J. Tromp (1998). Theoretical global seismology (2nd ed.). Princeton University Press. p. appendix C. ISBN 978-0-69100-1241.


  • D.O. Thompson; D.E. Chimenti (1997). Review of Progress in Quantitative Nondestructive Evaluation. Review Of Progress In Quantitative Nondestructive Evaluation. 16. Springer. p. 1708. ISBN 978-0-3064-55971.


  • H. Paetz; G. Schieck (2011). Nuclear Physics with Polarized Particles. Lecture Notes in Physics. 842. Springer. p. 31. ISBN 978-364-224-225-0.


  • V. Devanathan (1999). Angular Momentum Techniques in Quantum Mechanics. Fundamental Theories of Physics. 108. Springer. pp. 34, 61. ISBN 978-0-7923-5866-4.


  • V.D. Kleiman; R.N. Zare (1998). "5". A Companion to Angular Momentum. John Wiley & Sons. p. 112. ISBN 978-0-4711-9249-7.



Angular momentum and spin




  • Devanathan, V (2002). "Vectors and Tensors in Spherical Basis". Angular Momentum Techniques in Quantum Mechanics. Fundamental Theories of Physics. 108. pp. 24–33. doi:10.1007/0-306-47123-X_3. ISBN 978-0-306-47123-0.


  • K.T. Hecht (2000). Quantum mechanics. Graduate texts in contemporary physics. Springer. ISBN 978-0-387-989-198.



Condensed matter physics




  • J.A. Mettes; J.B. Keith; R.B. McClurg (2002). "Molecular Crystal Global Phase Diagrams:I Method of Construction" (PDF).


  • B.Henderson, R.H. Bartram (2005). Crystal-Field Engineering of Solid-State Laser Materials. Cambridge Studies in Modern Optics. 25. Cambridge University Press. p. 49. ISBN 978-0-52101-8012.


  • Edward U. Condon, and Halis Odabaşı (1980). Atomic Structure. CUP Archive. ISBN 978-0-5212-98933.


  • Melinda J. Duer, ed. (2008). "3". Solid State NMR Spectroscopy: Principles and Applications. John Wiley & Sons. p. 113. ISBN 978-0-4709-9938-7.


  • K.D. Bonin; V.V. Kresin (1997). "2". Electric - Dipole Polarizabilities of Atoms, Molecules and Clusters. World Scientific. pp. 14–15. ISBN 978-981-022-493-6.


  • A.E. McDermott, T.Polenova (2012). Solid State NMR Studies of Biopolymers. EMR handbooks. John Wiley & Sons. p. 42. ISBN 978-111-858-889-5.



Magnetic resonance




  • L.J. Mueller (2011). "Tensors and rotations in NMR". Concepts in Magnetic Resonance Part A. 38A (5): 221–235. doi:10.1002/cmr.a.20224.


  • M.S. Anwar (2004). "Spherical Tensor Operators in NMR" (PDF).


  • P. Callaghan (1993). Principles of nuclear magnetic resonance microscopy. Oxford University Press. pp. 56–57. ISBN 978-0-198-539-971.



Image processing




  • M. Reisert; H. Burkhardt (2009). S. Aja-Fernández, ed. Tensors in Image Processing and Computer Vision. Springer. ISBN 978-184-8822-993.


  • D.H. Laidlaw; J. Weickert (2009). Visualization and Processing of Tensor Fields: Advances and Perspectives. Mathematics and Visualization. Springer. ISBN 978-354-088-378-4.


  • M. Felsberg; E. Jonsson (2005). Energy Tensors: Quadratic, Phase Invariant Image Operators. Lecture Notes in Computer Science. 3663. Springer. pp. 493–500.


  • E. König; S. Kremer (1979). Tensor Operator Algebra for Point Groups. Lecture Notes in Computer Science. 3663. Springer. pp. 13–20.



External links



  • (2012) Clebsch-Gordon (sic) coefficients and the tensor spherical harmonics

  • The tensor spherical harmonics

  • (2010) Irreducible Tensor Operators and the Wigner-Eckart Theorem


  • Tensor operators[permanent dead link]

  • M. Fowler (2008), Tensor Operators

  • Tensor_Operators

  • (2009) Tensor Operators and the Wigner Eckart Theorem


  • The Wigner-Eckart theorem[permanent dead link]

  • (2004) Rotational Transformations and Spherical Tensor Operators

  • Tensor operators

  • Evaluation of the matrix elements for radiative transitions

  • D.K. Ghosh, (2013) Angular Momentum - III : Wigner- Eckart Theorem

  • B. Baragiola (2002) Tensor Operators

  • Spherical Tensors









這個網誌中的熱門文章

Hercules Kyvelos

Tangent Lines Diagram Along Smooth Curve

Yusuf al-Mu'taman ibn Hud