Symplectic matrix

In mathematics, a symplectic matrix is a 2n × 2n matrix M with real entries that satisfies the condition

 

 

 

 

(1)

where MT denotes the transpose of M and Ω is a fixed 2n × 2n nonsingular, skew-symmetric matrix. This definition can be extended to 2n × 2n matrices with entries in other fields, such as the complex numbers.

Typically Ω is chosen to be the block matrix

where In is the n × n identity matrix. The matrix Ω has determinant +1 and has an inverse given by Ω1 = ΩT = Ω.

Every symplectic matrix has determinant 1, and the 2n × 2n symplectic matrices with real entries form a subgroup of the special linear group SL(2n, R) under matrix multiplication. Topologically, this symplectic group is a connected noncompact real Lie group of real dimension n(2n + 1), and is denoted Sp(2n, R). The symplectic group can be defined as the set of linear transformations that preserve the symplectic form of a real symplectic vector space.

Examples of symplectic matrices include the identity matrix and the matrix .

Properties

Every symplectic matrix is invertible with the inverse matrix given by

Furthermore, the product of two symplectic matrices is, again, a symplectic matrix. This gives the set of all symplectic matrices the structure of a group. There exists a natural manifold structure on this group which makes it into a (real or complex) Lie group called the symplectic group.

It follows easily from the definition that the determinant of any symplectic matrix is ±1. Actually, it turns out that the determinant is always +1 for any field. One way to see this is through the use of the Pfaffian and the identity

Since and we have that det(M) = 1.

When the underlying field is real or complex, one can also show this by factoring the inequality .[1]

Suppose Ω is given in the standard form and let M be a 2n×2n block matrix given by

where A, B, C, D are n×n matrices. The condition for M to be symplectic is equivalent to the two following equivalent conditions[2]

symmetric, and
symmetric, and

When n = 1 these conditions reduce to the single condition det(M) = 1. Thus a 2×2 matrix is symplectic iff it has unit determinant.

With Ω in standard form, the inverse of M is given by

The group has dimension n(2n + 1). This can be seen by noting that is anti-symmetric. Since the space of anti-symmetric matrices has dimension , the identity imposes constraints on the coefficients of and leaves with n(2n+1) independent coefficients.

Symplectic transformations

In the abstract formulation of linear algebra, matrices are replaced with linear transformations of finite-dimensional vector spaces. The abstract analog of a symplectic matrix is a symplectic transformation of a symplectic vector space. Briefly, a symplectic vector space is a 2n-dimensional vector space V equipped with a nondegenerate, skew-symmetric bilinear form ω called the symplectic form.

A symplectic transformation is then a linear transformation L : VV which preserves ω, i.e.

Fixing a basis for V, ω can be written as a matrix Ω and L as a matrix M. The condition that L be a symplectic transformation is precisely the condition that M be a symplectic matrix:

Under a change of basis, represented by a matrix A, we have

One can always bring Ω to either the standard form given in the introduction or the block diagonal form described below by a suitable choice of A.

The matrix Ω

Symplectic matrices are defined relative to a fixed nonsingular, skew-symmetric matrix Ω. As explained in the previous section, Ω can be thought of as the coordinate representation of a nondegenerate skew-symmetric bilinear form. It is a basic result in linear algebra that any two such matrices differ from each other by a change of basis.

The most common alternative to the standard Ω given above is the block diagonal form

This choice differs from the previous one by a permutation of basis vectors.

Sometimes the notation J is used instead of Ω for the skew-symmetric matrix. This is a particularly unfortunate choice as it leads to confusion with the notion of a complex structure, which often has the same coordinate expression as Ω but represents a very different structure. A complex structure J is the coordinate representation of a linear transformation that squares to 1, whereas Ω is the coordinate representation of a nondegenerate skew-symmetric bilinear form. One could easily choose bases in which J is not skew-symmetric or Ω does not square to 1.

Given a hermitian structure on a vector space, J and Ω are related via

where is the metric. That J and Ω usually have the same coordinate expression (up to an overall sign) is simply a consequence of the fact that the metric g is usually the identity matrix.

Diagonalisation and decomposition

  • For any positive definite symmetric real symplectic matrix S there exists U in U(2n,R) such that

where the diagonal elements of D are the eigenvalues of S.[3]
  • Any real symplectic matrix can be decomposed as a product of three matrices:

 

 

 

 

(2)

such that O and O' are both symplectic and orthogonal and D is positive-definite and diagonal.[4] This decomposition is closely related to the singular value decomposition of a matrix and is known as an 'Euler' or 'Bloch-Messiah' decomposition.

Complex matrices

If instead M is a 2n×2n matrix with complex entries, the definition is not standard throughout the literature. Many authors [5] adjust the definition above to

 

 

 

 

(3)

where M* denotes the conjugate transpose of M. In this case, the determinant may not be 1, but will have absolute value 1. In the 2×2 case (n=1), M will be the product of a real symplectic matrix and a complex number of absolute value 1.

Other authors [6] retain the definition (1) for complex matrices and call matrices satisfying (3) conjugate symplectic.

Applications

Transformations described by symplectic matrices play an important role in quantum optics and in continuous-variable quantum information theory. For instance, symplectic matrices can be used to describe Gaussian (Bogoliubov) transformations of a quantum state of light.[7] In turn, the Bloch-Messiah decomposition (2) means that such an arbitrary Gaussian transformation can be represented as a set of two passive linear-optical interferometers (corresponding to orthogonal matrices O and O' ) intermitted by a layer of active non-linear squeezing transformations (given in terms of the matrix D).[8] In fact, one can circumvent the need for such in-line active squeezing transformations if two-mode squeezed vacuum states are available as a prior resource only.[9]

gollark: I mean, Islam has day-night-cycle-based prayers (problematic in space) and the pray-toward-the-Mecca requirement.
gollark: I wonder, does Judaism have possible issues translating to space like Islam?
gollark: The internet adopting it as some sort of meme thing and interjecting it into random unrelated conversations annoyed me enough that I don't really care.
gollark: I don't really know what happened with that, and honestly I don't care.
gollark: Although I think that performance on older devices getting worse is *generally* because of software developers getting used to having more power to throw at things, not a conspiracy.

See also

References

  1. Rim, Donsub (2017). "An elementary proof that symplectic matrices have determinant one". Adv. Dyn. Syst. Appl. 12 (1): 15–20. arXiv:1505.04240. Bibcode:2015arXiv150504240R.
  2. de Gosson, Maurice. "Introduction to Symplectic Mechanics: Lectures I-II-III" (PDF).
  3. de Gosson, Maurice A. (2011). Symplectic Methods in Harmonic Analysis and in Mathematical Physics - Springer. doi:10.1007/978-3-7643-9992-4. ISBN 978-3-7643-9991-7.
  4. Ferraro et. al. 2005 Section 1.3. ... Title?
  5. Xu, H. G. (July 15, 2003). "An SVD-like matrix decomposition and its applications". Linear Algebra and Its Applications. 368: 1–24. doi:10.1016/S0024-3795(03)00370-7. hdl:1808/374.
  6. Mackey, D. S.; Mackey, N. (2003). "On the Determinant of Symplectic Matrices". Numerical Analysis Report. 422. Manchester, England: Manchester Centre for Computational Mathematics. Cite journal requires |journal= (help)
  7. Weedbrook, Christian; Pirandola, Stefano; García-Patrón, Raúl; Cerf, Nicolas J.; Ralph, Timothy C.; Shapiro, Jeffrey H.; Lloyd, Seth (2012). "Gaussian quantum information". Reviews of Modern Physics. 84 (2): 621–669. arXiv:1110.3234. Bibcode:2012RvMP...84..621W. doi:10.1103/RevModPhys.84.621.
  8. Braunstein, Samuel L. (2005). "Squeezing as an irreducible resource". Physical Review A. 71 (5): 055801. arXiv:quant-ph/9904002. Bibcode:2005PhRvA..71e5801B. doi:10.1103/PhysRevA.71.055801.
  9. Chakhmakhchyan, Levon; Cerf, Nicolas (2018). "Simulating arbitrary Gaussian circuits with linear optics". Physical Review A. 98 (6): 062314. arXiv:1803.11534. Bibcode:2018PhRvA..98f2314C. doi:10.1103/PhysRevA.98.062314.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.