Tensor operator

Summary

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]

The general notion of scalar, vector, and tensor operators edit

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  , traveling with a definite center of mass momentum,  , in the   direction. If we rotate the system by   about the   axis, the momentum will change to  , which is in the   direction. The center-of-mass kinetic energy of the molecule will, however, be unchanged at  . 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   and  . 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,  , and the spin angular momentum,  . (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   of the two vector operators,   and  , 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

 
Here, the indices   and   can independently take on the values 1, 2, and 3 (or  ,  , and  ) corresponding to the three Cartesian axes, the index   runs over all particles (electrons and nuclei) in the molecule,   is the charge on particle  , and   is the  -th component of the position of this particle. Each term in the sum is a tensor operator. In particular, the nine products   together form a second rank tensor, formed by taking the outer product of the vector operator   with itself.

Rotations of quantum states edit

Quantum rotation operator edit

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

 

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

 

and let   be a rotation matrix. According to the Rodrigues' rotation formula, the rotation operator then amounts to

 

An operator   is invariant under a unitary transformation U if

 
in this case for the rotation  ,
 

Angular momentum eigenkets edit

The orthonormal basis set for total angular momentum is  , 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 within the j subspace

 

rotates to a new state by:

 

Using the completeness condition:

 

we have

 

Introducing the Wigner D matrix elements:

 

gives the matrix multiplication:

 

For one basis ket:

 

For the case of orbital angular momentum, the eigenstates   of the orbital angular momentum operator L and solutions of Laplace's equation on a 3d sphere are spherical harmonics:

 

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:

 

So a spherical harmonic can also be written  . Spherical harmonic states   rotate according to the inverse rotation matrix  , while   rotates by the initial rotation matrix  .

 

Rotation of tensor operators edit

We define the Rotation of an operator by requiring that the expectation value of the original operator   with respect to the initial state be equal to the expectation value of the rotated operator with respect to the rotated state,

 

Now as,

 

we have,

 

since,   is arbitrary,

 

Scalar operators edit

A scalar operator is invariant under rotations:[2]

 

This is equivalent to saying a scalar operator commutes with the rotation generators:

 

Examples of scalar operators include

  • the energy operator:
     
  • potential energy V (in the case of a central potential only)
     
  • kinetic energy T:
     
  • the spin–orbit coupling:
     

Vector operators edit

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

 
Any observable vector quantity of a quantum mechanical system should be invariant of the choice of frame of reference. The transformation of expectation value vector which applies for any wavefunction, ensures the above equality. In Dirac notation:
 
where the RHS is due to the rotation transformation acting on the vector formed by expectation values. Since |Ψ is any quantum state, the same result follows:
 
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.

From the above relation for infinitesimal rotations and the Baker Hausdorff lemma, by equating coefficients of order  , one can derive the commutation relation with the rotation generator:[2]

 
where εijk is the Levi-Civita symbol, which all vector operators must satisfy, by construction. The above commutator rule can also be used as an alternative definition for vector operators which can be shown by using the Baker Hausdorff lemma. As the symbol εijk is a pseudotensor, pseudovector operators are invariant up to a sign: +1 for proper rotations and −1 for improper rotations.

Since operators can be shown to form a vector operator by their commutation relation with angular momentum components (which are generators of rotation), its examples include:

  • the position operator:
     
  • the momentum operator:
     

and peusodovector operators include

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

Scalar operators from vector operators edit

If   and   are two vector operators, the dot product between the two vector operators can be defined as:

 

Under rotation of coordinates, the newly defined operator transforms as:

 
Rearranging terms and using transpose of rotation matrix as its inverse property:
 
Where the RHS is the   operator originally defined. Since the dot product defined is invariant under rotation transformation, it is said to be a scalar operator.

Spherical vector operators edit

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

 

using   the various commutators with the rotation generators and ladder operators are:
 

which are of similar form of

 


In the spherical basis, the generators of rotation are:

 

From the transformation of operators and Baker Hausdorff lemma:

 

compared to

 

it can be argued that the commutator with operator replaces the action of operator on state for transformations of operators as compared with that of states:

 

The rotation transformation in the spherical basis (originally written in the Cartesian basis) is then, due to similarity of commutation and operator shown above:

 

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

Tensor operators edit

In general, a tensor operator is one that transforms according to a tensor:

 
where the basis are transformed by   or the vector components transform by  .

In the subsequent discussion surrounding tensor operators, the index notation regarding covariant/contravariant behavior is ignored entirely. Instead, contravariant components is implied by context. Hence for an n times contravariant tensor:[2]

 

Examples of tensor operators edit

  • The Quadrupole moment operator,
     
  • Components of two tensor vector operators can be multiplied to give another Tensor operator.
     
    In general, n number of tensor operators will also give another tensor operator
     
    or,
     

Note: In general, a tensor operator cannot be written as the tensor product of other tensor operators as given in the above example.

Tensor operator from vector operators edit

If   and   are two three dimensional vector operators, then a rank 2 Cartesian dyadic tensors can be formed from nine operators of form  ,

 
Rearranging terms, we get:
 
The RHS of the equation is change of basis equation for twice contravariant tensors where the basis are transformed by   or the vector components transform by   which matches transformation of vector operator components. Hence the operator tensor described forms a rank 2 tensor, in tensor representation,
 
Similarly, an n-times contravariant tensor operator can be formed similarly by n vector operators.

We observe that the subspace spanned by linear combinations of the rank two tensor components form an invariant subspace, ie. the subspace does not change under rotation since the transformed components itself is a linear combination of the tensor components. However, this subspace is not irreducible ie. it can be further divided into invariant subspaces under rotation. Otherwise, the subspace is called reducible. In other words, there exists specific sets of different linear combinations of the components such that they transforms into a linear combination of the same set under rotation.[3] In the above example, we will show that the 9 independent tensor components can be divided into a set of 1, 3 and 5 combination of operators that each form irreducible invariant subspaces.

Irreducible tensor operators edit

The subspace spanned by   can be divided two subspaces; three independent antisymmetric components   and six independent symmetric component  , defined as   and  . Using the   transformation under rotation formula, it can be shown that both   and   are transformed into a linear combination of members of its own sets. Although   is irreducible, the same cannot be said about  .

The six independent symmetric component set can be divided into five independent traceless symmetric component and the invariant trace can be its own subspace.


Hence, the invariant subspaces of   are formed respectively by:

  1. One invariant trace of the tensor,  
  2. Three linearly independent antisymmetric components from:  
  3. Five linearly independent traceless symmetric components from  


If  , the invariant subspaces of   formed are represented by:[4]

  1. One invariant scalar operator  
  2. Three linearly independent components from  
  3. Five linearly independent components from  


From the above examples, the nine component   are split into subspaces formed by one, three and five components. These numbers add up to the number of components of the original tensor in a manner similar to the dimension of vector subspaces adding to the dimension of the space that is a direct sum of these subspaces. Similarly, every element of   can be expressed in terms of a linear combination of components from its invariant subspaces:

 

or

 

where:

 
 
 

In general cartesian tensors of rank greater than 1 are reducible. In quantum mechanics, this particular example bears resemblance to the addition of two spin one particles where both are 3 dimensional, hence the total space being 9 dimensional, can be formed by spin 0, spin 1 and spin 2 systems each having 1 dimensional, 3 dimensional and 5 dimensional space respectively.[4] 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. Each of the irreducible representations T(0), T(1), T(2) ... transform like angular momentum eigenstates according to the number of independent components.

It is possible that a given tensor may have one or more of these components vanish. For example, the quadrupole moment tensor is already symmetric and traceless, and hence has only 5 independent components to begin with.[3]

Spherical tensor operators edit

Spherical tensor operators are generally defined as operators with the following transformation rule, under rotation of coordinate system:

 


The commutation relations can be found by expanding LHS and RHS as:[4]

 

Simplifying and applying limits to select only first order terms, we get:

 

For choices of   or  , we get:

 
Note the similarity of the above to:
 
Since   and   are linear combinations of  , they share the same similarity due to linearity.


If, only the commutation relations hold, using the following relation,  

we find due to similarity of actions of   on wavefunction   and the commutation relations on  , that:

 

where the exponential form is given by Baker–Hausdorff lemma. Hence, the above commutation relations and the transformation property are equivalent definitions of spherical tensor operators. It can also be shown that   transform like a vector due to their commutation relation.


In the following section, construction of spherical tensors will be discussed. For example, since example of spherical vector operators is shown, it can be used to construct higher order spherical tensor operators. In general, spherical tensor operators can be constructed from two perspectives.[5] 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. Or by 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.[citation needed]

Construction using Clebsch–Gordan coefficients edit

Combination of two spherical tensors   and  in the following manner involving the Clebsch–Gordan coefficients can be proved to give another spherical tensor of the form:[4]

 

This equation can be used to construct higher order spherical tensor operators, for example, second order spherical tensor operators using two first order spherical tensor operators, say A and B, discussed previously:

 

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

 
and the finite rotation transformation in the spherical basis can be verified:
 

Using Spherical Harmonics edit

Define an operator by its spectrum:

 
Since for spherical harmonics under rotation:
 
It can also been shown that:
 
Then  , where   is a vector operator, also transforms in the same manner ie, is a spherical tensor operator. The process involves expressing   in terms of x, y and z and replacing x, y and z with operators Vx Vy and Vz which from vector operator. The resultant operator is hence a spherical tensor operator  .^ This may include constant due to normalization from spherical harmonics which is meaningless in context of operators.

The Hermitian adjoint of a spherical tensor may be defined as

 
There is some arbitrariness in the choice of the phase factor: any factor containing (−1)±q will satisfy the commutation relations.[6] The above choice of phase has the advantages of being real and that the tensor product of two commuting Hermitian operators is still Hermitian.[7] Some authors define it with a different sign on q, without the k, or use only the floor of k.[8]

Angular momentum and spherical harmonics edit

Orbital angular momentum and spherical harmonics edit

Orbital angular momentum operators have the ladder operators:

 

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 edit

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:

 

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

Applications edit

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) edit

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,

 

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

 

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,

 

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

 

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,

 

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,

 

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,

 

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 edit

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.[5]

Image processing and computer graphics edit

See also edit

References edit

Notes edit

  1. ^ Jeevanjee, Nadir (2015). An Introduction to Tensors and Group Theory for Physicists (2nd ed.). Birkhauser. ISBN 978-0-8176-4714-8.
  2. ^ a b c d e E. Abers (2004). "5". Quantum Mechanics. Addison Wesley. ISBN 978-0-13-146100-0.
  3. ^ a b Littlejohn, Robert G. (23 September 2023). "Irreducible Tensor Operators and the Wigner-Eckart Theorem" (PDF). Archived (PDF) from the original on 10 February 2023. Retrieved 23 September 2023.
  4. ^ a b c d Sakurai, Jun J.; Napolitano, Jim J. (2014). Modern quantum mechanics (2nd ed.). New Dehli: Pearson Education India. ISBN 978-93-325-1900-8.
  5. ^ a b R.D. Nielsen; B.H. Robinson (2006). "The Spherical Tensor Formalism Applied to Relaxation in Magnetic Resonance". Concepts in Magnetic Resonance Part A. 28A (4): 270–271. doi:10.1002/cmr.a.20055. Retrieved 2023-04-06.

Sources edit

  • 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. S2CID 8889942.
  • 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 edit

Spherical harmonics edit

  • 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. Vol. 16. Springer. p. 1708. ISBN 978-0-3064-55971.
  • H. Paetz; G. Schieck (2011). Nuclear Physics with Polarized Particles. Lecture Notes in Physics. Vol. 842. Springer. p. 31. ISBN 978-364-224-225-0.
  • V. Devanathan (1999). Angular Momentum Techniques in Quantum Mechanics. Fundamental Theories of Physics. Vol. 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 edit

  • Devanathan, V (2002). "Vectors and Tensors in Spherical Basis". Angular Momentum Techniques in Quantum Mechanics. Fundamental Theories of Physics. Vol. 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 edit

  • 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. Vol. 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 edit

  • 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. S2CID 8889942.
  • 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 edit

  • 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. Vol. 3663. Springer. pp. 493–500.
  • E. König; S. Kremer (1979). "Tensor Operator Algebra for Point Groups". Magnetism Diagrams for Transition Metal Ions. Lecture Notes in Computer Science. Vol. 3663. Springer. pp. 13–20. doi:10.1007/978-1-4613-3003-5_3. ISBN 978-1-4613-3005-9.

External links edit

  • (2012) Clebsch-Gordon (sic) coefficients and the tensor spherical harmonics
  • The tensor spherical harmonics
  • (2010) Irreducible Tensor Operators and the Wigner-Eckart Theorem Archived 2014-07-20 at the Wayback Machine
  • 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