Generalizations of Pauli matrices

In mathematics and physics, in particular quantum information, the term generalized Pauli matrices refers to families of matrices which generalize the (linear algebraic) properties of the Pauli matrices. Here, a few classes of such matrices are summarized.

Multi-qubit Pauli matrices (Hermitian) edit

This method of generalizing the Pauli matrices refers to a generalization from a single 2-level system (qubit) to multiple such systems. In particular, the generalized Pauli matrices for a group of   qubits is just the set of matrices generated by all possible products of Pauli matrices on any of the qubits.[1]

The vector space of a single qubit is   and the vector space of   qubits is  . We use the tensor product notation

 

to refer to the operator on   that acts as a Pauli matrix on the  th qubit and the identity on all other qubits. We can also use   for the identity, i.e., for any   we use  . Then the multi-qubit Pauli matrices are all matrices of the form

 ,

i.e., for   a vector of integers between 0 and 4. Thus there are   such generalized Pauli matrices if we include the identity   and   if we do not.

Higher spin matrices (Hermitian) edit

The traditional Pauli matrices are the matrix representation of the   Lie algebra generators  ,  , and   in the 2-dimensional irreducible representation of SU(2), corresponding to a spin-1/2 particle. These generate the Lie group SU(2).

For a general particle of spin  , one instead utilizes the  -dimensional irreducible representation.

Generalized Gell-Mann matrices (Hermitian) edit

This method of generalizing the Pauli matrices refers to a generalization from 2-level systems (Pauli matrices acting on qubits) to 3-level systems (Gell-Mann matrices acting on qutrits) and generic  -level systems (generalized Gell-Mann matrices acting on qudits).

Construction edit

Let   be the matrix with 1 in the jk-th entry and 0 elsewhere. Consider the space of   complex matrices,  , for a fixed  .

Define the following matrices,

 

and

 

The collection of matrices defined above without the identity matrix are called the generalized Gell-Mann matrices, in dimension  .[2][3] The symbol ⊕ (utilized in the Cartan subalgebra above) means matrix direct sum.

The generalized Gell-Mann matrices are Hermitian and traceless by construction, just like the Pauli matrices. One can also check that they are orthogonal in the Hilbert–Schmidt inner product on  . By dimension count, one sees that they span the vector space of   complex matrices,  . They then provide a Lie-algebra-generator basis acting on the fundamental representation of  .

In dimensions   = 2 and 3, the above construction recovers the Pauli and Gell-Mann matrices, respectively.

Sylvester's generalized Pauli matrices (non-Hermitian) edit

A particularly notable generalization of the Pauli matrices was constructed by James Joseph Sylvester in 1882.[4] These are known as "Weyl–Heisenberg matrices" as well as "generalized Pauli matrices".[5][6]

Framing edit

The Pauli matrices   and   satisfy the following:

 

The so-called Walsh–Hadamard conjugation matrix is

 

Like the Pauli matrices,   is both Hermitian and unitary.   and   satisfy the relation

 

The goal now is to extend the above to higher dimensions,  .

Construction: The clock and shift matrices edit

Fix the dimension   as before. Let  , a root of unity. Since   and  , the sum of all roots annuls:

 

Integer indices may then be cyclically identified mod d.

Now define, with Sylvester, the shift matrix

 

and the clock matrix,

 

These matrices generalize   and  , respectively.

Note that the unitarity and tracelessness of the two Pauli matrices is preserved, but not Hermiticity in dimensions higher than two. Since Pauli matrices describe quaternions, Sylvester dubbed the higher-dimensional analogs "nonions", "sedenions", etc.

These two matrices are also the cornerstone of quantum mechanical dynamics in finite-dimensional vector spaces[7][8][9] as formulated by Hermann Weyl, and they find routine applications in numerous areas of mathematical physics.[10] The clock matrix amounts to the exponential of position in a "clock" of   hours, and the shift matrix is just the translation operator in that cyclic vector space, so the exponential of the momentum. They are (finite-dimensional) representations of the corresponding elements of the Weyl-Heisenberg group on a  -dimensional Hilbert space.

The following relations echo and generalize those of the Pauli matrices:

 

and the braiding relation,

 

the Weyl formulation of the CCR, and can be rewritten as

 

On the other hand, to generalize the Walsh–Hadamard matrix  , note

 

Define, again with Sylvester, the following analog matrix,[11] still denoted by   in a slight abuse of notation,

 

It is evident that   is no longer Hermitian, but is still unitary. Direct calculation yields

 

which is the desired analog result. Thus,  , a Vandermonde matrix, arrays the eigenvectors of  , which has the same eigenvalues as  .

When  ,   is precisely the discrete Fourier transform matrix, converting position coordinates to momentum coordinates and vice versa.

Definition edit

The complete family of   unitary (but non-Hermitian) independent matrices   is defined as follows:

 

This provides Sylvester's well-known trace-orthogonal basis for  , known as "nonions"  , "sedenions"  , etc...[12][13]

This basis can be systematically connected to the above Hermitian basis.[14] (For instance, the powers of  , the Cartan subalgebra, map to linear combinations of the   matrices.) It can further be used to identify  , as  , with the algebra of Poisson brackets.

Properties edit

With respect to the Hilbert–Schmidt inner product on operators,  , Sylvester's generalized Pauli operators are orthogonal and normalized to  :

 .

This can be checked directly from the above definition of  .

See also edit

Notes edit

  1. ^ Brown, Adam R.; Susskind, Leonard (2018-04-25). "Second law of quantum complexity". Physical Review D. 97 (8): 086015. arXiv:1701.01107. Bibcode:2018PhRvD..97h6015B. doi:10.1103/PhysRevD.97.086015. S2CID 119199949.
  2. ^ Kimura, G. (2003). "The Bloch vector for N-level systems". Physics Letters A. 314 (5–6): 339–349. arXiv:quant-ph/0301152. Bibcode:2003PhLA..314..339K. doi:10.1016/S0375-9601(03)00941-1. S2CID 119063531.
  3. ^ Bertlmann, Reinhold A.; Philipp Krammer (2008-06-13). "Bloch vectors for qudits". Journal of Physics A: Mathematical and Theoretical. 41 (23): 235303. arXiv:0806.1174. Bibcode:2008JPhA...41w5303B. doi:10.1088/1751-8113/41/23/235303. ISSN 1751-8121. S2CID 118603188.
  4. ^ Sylvester, J. J., (1882), Johns Hopkins University Circulars I: 241-242; ibid II (1883) 46; ibid III (1884) 7–9. Summarized in The Collected Mathematics Papers of James Joseph Sylvester (Cambridge University Press, 1909) v III . online and further.
  5. ^ Appleby, D. M. (May 2005). "Symmetric informationally complete–positive operator valued measures and the extended Clifford group". Journal of Mathematical Physics. 46 (5): 052107. arXiv:quant-ph/0412001. Bibcode:2005JMP....46e2107A. doi:10.1063/1.1896384. ISSN 0022-2488.
  6. ^ Howard, Mark; Vala, Jiri (2012-08-15). "Qudit versions of the qubit π / 8 gate". Physical Review A. 86 (2): 022316. arXiv:1206.1598. Bibcode:2012PhRvA..86b2316H. doi:10.1103/PhysRevA.86.022316. ISSN 1050-2947. S2CID 56324846.
  7. ^ Weyl, H., "Quantenmechanik und Gruppentheorie", Zeitschrift für Physik, 46 (1927) pp. 1–46, doi:10.1007/BF02055756.
  8. ^ Weyl, H., The Theory of Groups and Quantum Mechanics (Dover, New York, 1931)
  9. ^ Santhanam, T. S.; Tekumalla, A. R. (1976). "Quantum mechanics in finite dimensions". Foundations of Physics. 6 (5): 583. Bibcode:1976FoPh....6..583S. doi:10.1007/BF00715110. S2CID 119936801.
  10. ^ For a serviceable review, see Vourdas A. (2004), "Quantum systems with finite Hilbert space", Rep. Prog. Phys. 67 267. doi:10.1088/0034-4885/67/3/R03.
  11. ^ Sylvester, J.J. (1867). "Thoughts on inverse orthogonal matrices, simultaneous sign-successions, and tessellated pavements in two or more colours, with applications to Newton's rule, ornamental tile-work, and the theory of numbers". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 34 (232): 461–475. doi:10.1080/14786446708639914.
  12. ^ Patera, J.; Zassenhaus, H. (1988). "The Pauli matrices in n dimensions and finest gradings of simple Lie algebras of type An−1". Journal of Mathematical Physics. 29 (3): 665. Bibcode:1988JMP....29..665P. doi:10.1063/1.528006.
  13. ^ Since all indices are defined cyclically mod d,  .
  14. ^ Fairlie, D. B.; Fletcher, P.; Zachos, C. K. (1990). "Infinite-dimensional algebras and a trigonometric basis for the classical Lie algebras". Journal of Mathematical Physics. 31 (5): 1088. Bibcode:1990JMP....31.1088F. doi:10.1063/1.528788.