III.88 Symplectic Manifolds

Gabriel P. Paternain


Symplectic geometry is the geometry that governs classical physics, and more generally plays an important role in helping us to understand the actions of groups on manifolds. It shares some features with Riemannian geometry and complex geometry, and there is an important special class of manifolds, the Kähler manifolds, in which all three geometric structures are unified.

1 Symplectic Linear Algebra

Just as RIEMANNIAN GEOMETRY [I.3 §6.10] is based on EUCLIDEAN GEOMETRY [I.3 §6.2], symplectic geometry is based on the geometry of the so-called linear symplectic space (Image2n, ω0).

Given two vectors v = (q, p) and v′ = (q′ , p′) in the plane Image2, the signed area ω0(v , v′) of the parallelogram spanned by v and v′ is given by the formula

Image

It can also be written using matrices and inner products as ω0(v, v′) = v′ · Jv, where J is the 2 x 2 matrix

Image

If a linear transformation A : Image2Image2 is area preserving and orientation preserving, then ω0(,Aυ′) = ωo(υ, υ′) for every υ and υ′.

Symplectic geometry studies two-dimensional signed area measurements like this, as well as transformations that preserve these measurements, but it applies to general spaces of dimension 2n rather than just to the plane.

If we split Image2n up as Imagen x Imagen, then we can write a vector υ in Image2n as v = (q, p), where q and p each belong to Imagen. The standard symplectic form ω0 : Image2n x Image2nImage is defined by the formula

ω0(υ,υ′) = p · q′ - q · p′,

where “·” denotes the usual inner product in Imagen. Geometrically, ω0(υ, υ′) can be interpreted as the sum of the signed areas of the parallelograms spanned by the projections of υ and υ′ to the qipi-planes. In terms of matrices, we can write

Image

 

where J is the 2n × 2n matrix

Image

and I is the n × n identity matrix.

A linear map A : Image2nImage2n that preserves the product ω0 of any two vectors (that is, ω0(Av,Av′) = ω0(v, v′) for all v, v′Image2n) is called a symplectic linear transformation; equivalently, a 2n × 2n matrix A is symplectic if and only if ATJA = J, where AT is the transpose of A. Symplectic linear transformations are to symplectic geometry as rigid motions are to Euclidean geometry. The set of all symplectic linear transformations of (Image2n, ω0) is one of the classical LIE GROUPS [III.48 §1] and is denoted by Sp(2n). One can show that symplectic matrices A ∈ Sp(2n) always have DETERMINANT [III.15] 1, and are thus volume preserving. However, the converse does not hold when n Image 2. For instance, if n = 2, the linear map

(q1, q2, p1, p2) → (aq1, q2/a, ap1, p2/a)

has determinant 1 for any a ≠ 0, but it is symplectic only if a2 = 1.

The standard symplectic form ω0 has three properties worth noting. First, it is bilinear: the expression ω0(υ,υ′) varies linearly in υv′ when υ′ is held fixed, and vice versa. Second, it is crntisymmetric: we have ω0(υ, υ′) = -(υ′, υ) for all υ and υ′, and in particular ω0(υ, υ) = 0. Finally, it is nondegenerate, which means that for every nonzero υ there is a nonzero υ′ such that ω0(υ, υ′) ≠ 0. The standard symplectic form ω0 is not the only form that obeys these three properties; however, it turns out that any form with these three properties can be converted into the standard form ω0 after an invertible linear change of variables. (This is a special case of Dgrboux’s theorem.) Thus (Image2n, ω0) is essentially the “only” linear symplectic geometry in 2n dimensions. There are no symplectic forms in odd-dimensional spaces.

2 Symplectic Diffeomorphisms of (Image2n, ω0)

In Euclidean geometry, all rigid motions are automatically linear (or affine) transformations. However, in symplectic geometry there are many more symplectic maps than just the symplectic linear transformations. These nonlinear symplectic maps in (Image2n, ω0) are one of the principal objects of study in symplectic geometry.

Let UImage2n be an open set. Recall that a map Image : UImage2n is called smooth if it has continuous partial derivatives of all orders. A diffeomorphism is a smooth map with smooth inverse.

A smooth nonlinear map Image : UImage2n is said to be symplectic if, for every xU, the Jacobian matrix Image (x) of first derivatives of Image is a symplectic linear transformation. Informally, a symplectic map is one that behaves like a symplectic linear transformation at infinitesimally small scales. Since symplectic linear transformations have determinant 1, we can conclude using several-variable calculus that a symplectic map is always locally volume preserving and locally invertible; roughly speaking, this means that the map Image : A →, Image(A) is invertible whenever A is a sufficiently small subset of U, and Image(A) has the same volume as A. However, the converse is not true when n Image 2; the class of symplectic maps is much more restricted than that of volume-preserving maps. In fact, Gromov’s non-squeezing theorem (see below) shows how striking this difference can be.

Symplectic maps have been around for quite a long time in Hamiltonian mechanics under the name of canonical transformations. We briefly explain this in the next subsection.

2.1 Hamilton’s Equations

How can we produce nonlinear symplectic maps? Let us begin by exploring a familiar example. Consider the motion of a simple pendulum with length l and mass m and let q(t)be the angle it makes with the vertical at time t. The equation of motion is

Image

where g is the acceleration due to gravity. If we define the momentum Image as Image = ml2 Image, then we may transform this second-order differential equation into a first-order system in the phase plane Image2, namely

Image

where the vector field X : Image2Image2 is given by the formula X(q,p) = (p/ml2, -mglsin q). For each (q(0), p(0)) Image Image2 there is a unique solution (q(t), p(t)) to (3) with initial condition (q(0), p(0)). Then for any fixed time t we obtain an evolution map (or flow) Imaget : Image2Image2 given by Imaget(q(0), p(0)) = (q(t), p(t)), which has the remarkable property of being area preserving. This can be deduced from the observation that X is divergence free, or in other words that

Image

In fact, for every time t, Image t is a symplectic map on (Image2, ω0).

More generally, any system in classical mechanics with finitely many degrees of freedom can be reformulated in a similar fashion, so that the evolution maps Imaget are always symplectic maps; in this context, they are also known as canonical transformations. The Irish mathematician WILLIAM ROWAN HAMILTON [VI.37] showed us how to do this in general more than 170 years ago. Given any smooth function H :Image2nImage (called the Hamiltonian), the system of first-order differential equations given by

Image

Image

will (under some mild growth assumptions on H, which we ignore here) give rise to evolution operators Imaget : Image2nImage2n, which are symplectic maps on (Image2n, ω0) for every time t. To see the connection with the symplectic form ω0, observe that we may rewrite (4) and (5) in the following equivalent form:

Image

where ∇ H is the usual GRADIENT [I.3§5.3] of H and J was defined in (2). From (6), (1), and the antisymmetry property of ω0, it is then not difficult to verify that Imaget is symplectic for every t (the main trick is to compute the derivative of ω0 (Image(x)v,Image(x)v′) n t and check that it equals zero).

We have already pointed out that symplectic maps are volume preserving. The preservation of volume by Hamiltonian systems (a result known as Liouville’s theorem) attracted considerable attention in the nineteenth century and it was a driving force in the development of ERGODIC THEORY [V.9], which studies recurrence properties of measure-preserving transformations.

Symplectic maps or canonical transformations play an important role in classical physics, as they allow one to replace a complicated system by an equivalent system that is simpler to analyze.

2.2 Gromov’s Nonsqueezing Theorem

What is the difference between a symplectic map and a volume-preserving map? In order to answer this question, suppose that we have two connected open sets U and V in Image2n and that we wish to embed one into the other using a symplectic map. This means that we are looking for a symplectic map Image : UV such that Image is a homeomorphism onto its image. We know such a Image must be volume preserving, so we clearly have the restriction that the volume of U should be at most the volume of V, but is this restriction all that matters? Consider the open ball B(R) = {x Image Image2n : |x| < R}, which has radius R and center at the origin, and clearly has finite volume. It is not hard to embed it symplectically into the infinite-volume cylinder given by

C(r) = {(q, p) Image Image2n : Image + Image < r2},

whatever the values of R and r. Indeed, the linear symplectic map

(q,p) Image (aq1, aq2, q3, . . . , qn, p1/a,p2/a,p3, . . . , pn)

will do the trick when a is sufficiently small and positive. However, the situation is radically different if instead we consider the infinite-volume cylinder

Z(r) = {(q, p) Image Image2n : Image + Image <r2}.

We could try with a similar linear map like

(q,p) Image (aqb1,q2/a,q3, . . . ,qn,ap1,p2/a,p3, . . . ,pn).

This map is volume preserving (it has determinant 1) and for a small it embeds B(R) into Z(r). However, it is symplectic only if a = 1, so it will give a symplectic embedding only if Rr. One is tempted to think that if R > r, then there should still be a nonlinear symplectic embedding squeezing B(R) into Z(r), but a remarkable theorem of Gromov from 1985 asserts that it is not possible to find such a map.

In spite of this deep result of Gromov, and other results that followed it, we still do not know much about how sets in Image2n embed into one another.

3 Synrplectic Manifolds

Recall from DIFFERENTIAL TOPOLOGY [IV.7] that a manifold of dimension d is a TOPOLOGICAL SPACE [III.90] such that each point has a neighborhood that is homeomorphic to an open set in Euclidean space Imaged. One can think of Imaged as a local model for this manifold, in the sense that it describes what the manifold looks like at very small distance scales. Recall also that a smooth manifold is one for which the “transition functions” are smooth. This means that if ψ : UImaged and φ : V → Imaged are coordinate charts, then the transition function ψ Image φ-1 between the open sets φ (UV) and ψ(UV) is smooth.

A symplectic manifold is defined similarly, but now the local model is the linear symplectic space (Image2n, ω0). More precisely, a symplectic manifold M is a manifold of dimension 2n that can be covered with domains of coordinate charts whose transition functions are symplectic diffenmorphisms of (Image2n, ω0).

Of course, any open set in (Image2n, ω0) is a symplectic manifold. An example of a compact symplectic manifold is the torus Image2n, which is obtained as the quotient of Image2n by the action of Image2n. In other words, two points x,y Image Image2n are equivalent if x - y has integer coordinates. Other important examples of symplectic manifolds include RIEMANN SURFACES [III.79], complex PROJECTIVE SPACE [III.72], and cotangent BUNDLES [IV.6 § 5]. However, it is a wide open problem to determine, given a compact manifold, whether it can be assigned a system of coordinate charts that makes it symplectic.

We have seen that in (Image2n, ω0), one can assign an Image2n “area” ω0 (v, v′) to any parallelogram in the space Image2n. In a symplectic manifold M, one can similarly assign an area ωp (v v′), but only to infinitesimal parallelograms based at a point p Image M. The axes of such a parallelogram are two infinitesimal vectors (or more precisely tangent vectors) υ and υ′ . There is a unique way to do this so that all the coordinate charts for M are symplectic diffeomorphisms. In the language of DIFFERENTIAL FORMS [III.16], the map Image Image ωImage is an antisymmetric nondegenerate 2-form, which can then be used to compute the “area” ∫s ω of nnrinfintesima1 two-dimensional surfaces S in M. One can show that for any sufficiently small closed surface S, the integral ∫sω vanishes, so ω is a closed form. Indeed, one can define a symplectic manifold more abstractly (without reference to charts) as a smooth manifold equipped with a closed, antisymmetric nondegenerate 2-form ω; a classical theorem of Darboux asserts that this abstract definition is equivalent to the more concrete definition using coordinate charts.

Finally, a special class of symplectic manifolds is given by Kähler manifolds. These are symplectic manifolds that are also complex manifolds, in such a way that the two structures are naturally compatible, a condition that generalizes the relationship (1). Observe that if one identifies points (q, p) in Image2n with points p + iq in Imagen, then the linear transformation J : Image2nImage2n becomes the operation of multiplication by i:

J: (z1 , . . . , zn) Image (iz1 , . . . , izn).

Thus the identity (1) relates the symplectic structure (as given by ω0), the complex structure (as given by J), and the Riemannian structure (as given by the dot product “·“). A complex manifold is a manifold that at small distance scales looks like regions of Imagen, with the transition functions required to be HOLQMORPWC [I.3 §5.6]. (A smooth map f : UImagenImagen is said to be holomorphic if each coordinate component of f (z1, , . . . , zn) is holomorphic in each variable zk.) On a complex manifold we can multiply tangent vectors by i. This gives us at each point p Image M a linear map Jp such that Imagev = -v for all tangent vectors v at p. A Kähler manifold is a complex manifold M with a symplectic structure ω (which computes signed areas of infinitesimal parallelograms) and a Riemannian metric g (which computes an inner product gImage(υ, υ′) of any two tangent vectors υ, υ′ at p); these two structures are linked together by the analogue of (1), namely

ωImage (υ, υ′) = gp(υ′ ,JImageυ).

Examples of Kähler manifolds include complex vector spaces Imagen, Riemann surfaces, and complex projective spaces ImageImagen.

An example of a compact symplectic manifold that is not Kähler can be obtained by taking the quotient of Image4 by a symplectic action of a group that looks like Image4 but with a group operation that differs from the usual one. The change in the group structure manifests itself as a topological property (an odd first Betti number) that prevents the quotient being Kähler.

Further Reading

Arnold, V. I. 1989. Mathematical Methods of Classical Mechanics, 2nd edn. Graduate Texts in Mathematics, volume 60. New York: Springer.

McDuff, D., and D. Salamon. 1998. Introduction to Symplectic Topology, 2nd edn. Oxford Mathematical Monographs. Oxford: Clarendon Press/Oxford University Press.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset