Oscillator representation

From Wikipedia the free encyclopedia

In mathematics, the oscillator representation is a projective unitary representation of the symplectic group, first investigated by Irving Segal, David Shale, and André Weil. A natural extension of the representation leads to a semigroup of contraction operators, introduced as the oscillator semigroup by Roger Howe in 1988. The semigroup had previously been studied by other mathematicians and physicists, most notably Felix Berezin in the 1960s. The simplest example in one dimension is given by SU(1,1). It acts as Möbius transformations on the extended complex plane, leaving the unit circle invariant. In that case the oscillator representation is a unitary representation of a double cover of SU(1,1) and the oscillator semigroup corresponds to a representation by contraction operators of the semigroup in SL(2,C) corresponding to Möbius transformations that take the unit disk into itself.

The contraction operators, determined only up to a sign, have kernels that are Gaussian functions. On an infinitesimal level the semigroup is described by a cone in the Lie algebra of SU(1,1) that can be identified with a light cone. The same framework generalizes to the symplectic group in higher dimensions, including its analogue in infinite dimensions. This article explains the theory for SU(1,1) in detail and summarizes how the theory can be extended.

Historical overview[edit]

The mathematical formulation of quantum mechanics by Werner Heisenberg and Erwin Schrödinger was originally in terms of unbounded self-adjoint operators on a Hilbert space. The fundamental operators corresponding to position and momentum satisfy the Heisenberg commutation relations. Quadratic polynomials in these operators, which include the harmonic oscillator, are also closed under taking commutators.

A large amount of operator theory was developed in the 1920s and 1930s to provide a rigorous foundation for quantum mechanics. Part of the theory was formulated in terms of unitary groups of operators, largely through the contributions of Hermann Weyl, Marshall Stone and John von Neumann. In turn these results in mathematical physics were subsumed within mathematical analysis, starting with the 1933 lecture notes of Norbert Wiener, who used the heat kernel for the harmonic oscillator to derive the properties of the Fourier transform.

The uniqueness of the Heisenberg commutation relations, as formulated in the Stone–von Neumann theorem, was later interpreted within group representation theory, in particular the theory of induced representations initiated by George Mackey. The quadratic operators were understood in terms of a projective unitary representation of the group SU(1,1) and its Lie algebra. Irving Segal and David Shale generalized this construction to the symplectic group in finite and infinite dimensions—in physics, this is often referred to as bosonic quantization: it is constructed as the symmetric algebra of an infinite-dimensional space. Segal and Shale have also treated the case of fermionic quantization, which is constructed as the exterior algebra of an infinite-dimensional Hilbert space. In the special case of conformal field theory in 1+1 dimensions, the two versions become equivalent via the so-called "boson-fermion correspondence." Not only does this apply in analysis where there are unitary operators between bosonic and fermionic Hilbert spaces, but also in the mathematical theory of vertex operator algebras. Vertex operators themselves originally arose in the late 1960s in theoretical physics, particularly in string theory.

André Weil later extended the construction to p-adic Lie groups, showing how the ideas could be applied in number theory, in particular to give a group theoretic explanation of theta functions and quadratic reciprocity. Several physicists and mathematicians observed the heat kernel operators corresponding to the harmonic oscillator were associated to a complexification of SU(1,1): this was not the whole of SL(2,C), but instead a complex semigroup defined by a natural geometric condition. The representation theory of this semigroup, and its generalizations in finite and infinite dimensions, has applications both in mathematics and theoretical physics.[1]

Semigroups in SL(2,C)[edit]

The group:

is a subgroup of Gc = SL(2,C), the group of complex 2 × 2 matrices with determinant 1. If G1 = SL(2,R) then

This follows since the corresponding Möbius transformation is the Cayley transform which carries the upper half plane onto the unit disk and the real line onto the unit circle.

The group SL(2,R) is generated as an abstract group by

and the subgroup of lower triangular matrices

Indeed, the orbit of the vector

under the subgroup generated by these matrices is easily seen to be the whole of R2 and the stabilizer of v in G1 lies in inside this subgroup.

The Lie algebra of SU(1,1) consists of matrices

The period 2 automorphism σ of Gc

with

has fixed point subgroup G since

Similarly the same formula defines a period two automorphism σ of the Lie algebra of Gc, the complex matrices with trace zero. A standard basis of over C is given by

Thus for −1 ≤ m, n ≤ 1

There is a direct sum decomposition

where is the +1 eigenspace of σ and the –1 eigenspace.

The matrices X in have the form

Note that

The cone C in is defined by two conditions. The first is By definition this condition is preserved under conjugation by G. Since G is connected it leaves the two components with x > 0 and x < 0 invariant. The second condition is

The group Gc acts by Möbius transformations on the extended complex plane. The subgroup G acts as automorphisms of the unit disk D. A semigroup H of Gc, first considered by Olshanskii (1981), can be defined by the geometric condition:

The semigroup can be described explicitly in terms of the cone C:[2]

In fact the matrix X can be conjugated by an element of G to the matrix

with

Since the Möbius transformation corresponding to exp Y sends z to e−2yz, it follows that the right hand side lies in the semigroup. Conversely if g lies in H it carries the closed unit disk onto a smaller closed disk in its interior. Conjugating by an element of G, the smaller disk can be taken to have centre 0. But then for appropriate y, the element carries D onto itself so lies in G.

A similar argument shows that the closure of H, also a semigroup, is given by

From the above statement on conjugacy, it follows that

where

If

then

since the latter is obtained by taking the transpose and conjugating by the diagonal matrix with entries ±1. Hence H also contains

which gives the inverse matrix if the original matrix lies in SU(1,1).

A further result on conjugacy follows by noting that every element of H must fix a point in D, which by conjugation with an element of G can be taken to be 0. Then the element of H has the form

The set of such lower triangular matrices forms a subsemigroup H0 of H.

Since

every matrix in H0 is conjugate to a diagonal matrix by a matrix M in H0.

Similarly every one-parameter semigroup S(t) in H fixes the same point in D so is conjugate by an element of G to a one-parameter semigroup in H0.

It follows that there is a matrix M in H0 such that

with S0(t) diagonal. Similarly there is a matrix N in H0 such that

The semigroup H0 generates the subgroup L of complex lower triangular matrices with determinant 1 (given by the above formula with a ≠ 0). Its Lie algebra consists of matrices of the form

In particular the one parameter semigroup exp tZ lies in H0 for all t > 0 if and only if and

This follows from the criterion for H or directly from the formula

The exponential map is known not to be surjective in this case, even though it is surjective on the whole group L. This follows because the squaring operation is not surjective in H. Indeed, since the square of an element fixes 0 only if the original element fixes 0, it suffices to prove this in H0. Take α with |α| < 1 and

If a = α2 and

with

then the matrix

has no square root in H0. For a square root would have the form

On the other hand,

The closed semigroup is maximal in SL(2,C): any larger semigroup must be the whole of SL(2,C).[3][4][5][6][7]

Using computations motivated by theoretical physics, Ferrara et al. (1973) introduced the semigroup , defined through a set of inequalities. Without identification as a compression semigroup, they established the maximality of . Using the definition as a compression semigroup, maximality reduces to checking what happens when adding a new fractional transformation to . The idea of the proof depends on considering the positions of the two discs and . In the key cases, either one disc contains the other or they are disjoint. In the simplest cases, is the inverse of a scaling transformation or . In either case and generate an open neighbourhood of 1 and hence the whole of SL(2,C)

Later Lawson (1998) gave another more direct way to prove maximality by first showing that there is a g in S sending D onto the disk Dc, |z| > 1. In fact if then there is a small disk D1 in D such that xD1 lies in Dc. Then for some h in H, D1 = hD. Similarly yxD1 = Dc for some y in H. So g = yxh lies in S and sends D onto Dc. It follows that g2 fixes the unit disc D so lies in SU(1,1). So g−1 lies in S. If t lies in H then tgD contains gD. Hence So t−1 lies in S and therefore S contains an open neighbourhood of 1. Hence S = SL(2,C).

Exactly the same argument works for Möbius transformations on Rn and the open semigroup taking the closed unit sphere ||x|| ≤ 1 into the open unit sphere ||x|| < 1. The closure is a maximal proper semigroup in the group of all Möbius transformations. When n = 1, the closure corresponds to Möbius transformations of the real line taking the closed interval [–1,1] into itself.[8]

The semigroup H and its closure have a further piece of structure inherited from G, namely inversion on G extends to an antiautomorphism of H and its closure, which fixes the elements in exp C and its closure. For

the antiautomorphism is given by

and extends to an antiautomorphism of SL(2,C).

Similarly the antiautomorphism

leaves G1 invariant and fixes the elements in exp C1 and its closure, so it has analogous properties for the semigroup in G1.

Commutation relations of Heisenberg and Weyl[edit]

Let be the space of Schwartz functions on R. It is dense in the Hilbert space L2(R) of square-integrable functions on R. Following the terminology of quantum mechanics, the "momentum" operator P and "position" operator Q are defined on by

There operators satisfy the Heisenberg commutation relation

Both P and Q are self-adjoint for the inner product on inherited from L2(R).

Two one parameter unitary groups U(s) and V(t) can be defined on and L2(R) by

By definition

for , so that formally

It is immediate from the definition that the one parameter groups U and V satisfy the Weyl commutation relation

The realization of U and V on L2(R) is called the Schrödinger representation.

Fourier transform[edit]

The Fourier transform is defined on by[9]

It defines a continuous map of into itself for its natural topology.

Contour integration shows that the function

is its own Fourier transform.

On the other hand, integrating by parts or differentiating under the integral,

It follows that the operator on defined by

commutes with both Q (and P). On the other hand,

and since

lies in , it follows that

and hence

This implies the Fourier inversion formula:

and shows that the Fourier transform is an isomorphism of onto itself.

By Fubini's theorem

When combined with the inversion formula this implies that the Fourier transform preserves the inner product

so defines an isometry of onto itself.

By density it extends to a unitary operator on L2(R), as asserted by Plancherel's theorem.

Stone–von Neumann theorem[edit]

Suppose U(s) and V(t) are one parameter unitary groups on a Hilbert space satisfying the Weyl commutation relations

For let[10][11]

and define a bounded operator on by

Then

where

The operators T(F) have an important non-degeneracy property: the linear span of all vectors T(F)ξ is dense in .

Indeed, if fds and gdt define probability measures with compact support, then the smeared operators

satisfy

and converge in the strong operator topology to the identity operator if the supports of the measures decrease to 0.

Since U(f)V(g) has the form T(F), non-degeneracy follows.

When is the Schrödinger representation on L2(R), the operator T(F) is given by

It follows from this formula that U and V jointly act irreducibly on the Schrödinger representation since this is true for the operators given by kernels that are Schwartz functions. A concrete description is provided by Linear canonical transformations.

Conversely given a representation of the Weyl commutation relations on , it gives rise to a non-degenerate representation of the *-algebra of kernel operators. But all such representations are on an orthogonal direct sum of copies of L2(R) with the action on each copy as above. This is a straightforward generalisation of the elementary fact that the representations of the N × N matrices are on direct sums of the standard representation on CN. The proof using matrix units works equally well in infinite dimensions.

The one parameter unitary groups U and V leave each component invariant, inducing the standard action on the Schrödinger representation.

In particular this implies the Stone–von Neumann theorem: the Schrödinger representation is the unique irreducible representation of the Weyl commutation relations on a Hilbert space.

Oscillator representation of SL(2,R)[edit]

Given U and V satisfying the Weyl commutation relations, define

Then

so that W defines a projective unitary representation of R2 with cocycle given by

where and B is the symplectic form on R2 given by

By the Stone–von Neumann theorem, there is a unique irreducible representation corresponding to this cocycle.

It follows that if g is an automorphism of R2 preserving the form B, i.e. an element of SL(2,R), then there is a unitary π(g) on L2(R) satisfying the covariance relation

By Schur's lemma the unitary π(g) is unique up to multiplication by a scalar ζ with |ζ| = 1, so that π defines a projective unitary representation of SL(2,R).

This can be established directly using only the irreducibility of the Schrödinger representation. Irreducibility was a direct consequence of the fact the operators

with K a Schwartz function correspond exactly to operators given by kernels with Schwartz functions.

These are dense in the space of Hilbert–Schmidt operators, which, since it contains the finite rank operators, acts irreducibly.

The existence of π can be proved using only the irreducibility of the Schrödinger representation. The operators are unique up to a sign with

so that the 2-cocycle for the projective representation of SL(2,R) takes values ±1.

In fact the group SL(2,R) is generated by matrices of the form

and it can be verified directly that the following operators satisfy the covariance relations above:

The generators gi satisfy the following Bruhat relations, which uniquely specify the group SL(2,R):[12]

It can be verified by direct calculation that these relations are satisfied up to a sign by the corresponding operators, which establishes that the cocycle takes values ±1.

There is a more conceptual explanation using an explicit construction of the metaplectic group as a double cover of SL(2,R).[13] SL(2,R) acts by Möbius transformations on the upper half plane H. Moreover, if

then

The function

satisfies the 1-cocycle relation

For each g, the function m(g,z) is non-vanishing on H and therefore has two possible holomorphic square roots. The metaplectic group is defined as the group

By definition it is a double cover of SL(2,R) and is connected. Multiplication is given by

where

Thus for an element g of the metaplectic group there is a uniquely determined function m(g,z)1/2 satisfying the 1-cocycle relation.

If , then

lies in L2 and is called a coherent state.

These functions lie in a single orbit of SL(2,R) generated by

since for g in SL(2,R)

More specifically if g lies in Mp(2,R) then

Indeed, if this holds for g and h, it also holds for their product. On the other hand, the formula is easily checked if gt has the form gi and these are generators.

This defines an ordinary unitary representation of the metaplectic group.

The element (1,–1) acts as multiplication by –1 on L2(R), from which it follows that the cocycle on SL(2,R) takes only values ±1.

Maslov index[edit]

As explained in Lion & Vergne (1980), the 2-cocycle on SL(2,R) associated with the metaplectic representation, taking values ±1, is determined by the Maslov index.

Given three non-zero vectors u, v, w in the plane, their Maslov index is defined as the signature of the quadratic form on R3 defined by

Properties of the Maslov index:

  • it depends on the one-dimensional subspaces spanned by the vectors
  • it is invariant under SL(2,R)
  • it is alternating in its arguments, i.e. its sign changes if two of the arguments are interchanged
  • it vanishes if two of the subspaces coincide
  • it takes the values –1, 0 and +1: if u and v satisfy B(u,v) = 1 and w = au + bv, then the Maslov index is zero is if ab = 0 and is otherwise equal to minus the sign of ab

Picking a non-zero vector u0, it follows that the function

defines a 2-cocycle on SL(2,R) with values in the eighth roots of unity.

A modification of the 2-cocycle can be used to define a 2-cocycle with values in ±1 connected with the metaplectic cocycle.[14]

In fact given non-zero vectors u, v in the plane, define f(u,v) to be

  • i times the sign of B(u,v) if u and v are not proportional
  • the sign of λ if u = λv.

If

then

The representatives π(g) in the metaplectic representation can be chosen so that

where the 2-cocycle ω is given by

with

Holomorphic Fock space[edit]

Holomorphic Fock space (also known as the Segal–Bargmann space) is defined to be the vector space of holomorphic functions f(z) on C with

finite. It has inner product

is a Hilbert space with orthonormal basis

Moreover, the power series expansion of a holomorphic function in gives its expansion with respect to this basis.[15] Thus for z in C

so that evaluation at z is gives a continuous linear functional on In fact

where[16]

Thus in particular is a reproducing kernel Hilbert space.

For f in and z in C define

Then

so this gives a unitary representation of the Weyl commutation relations.[17] Now

It follows that the representation is irreducible.

Indeed, any function orthogonal to all the Ea must vanish, so that their linear span is dense in .

If P is an orthogonal projection commuting with W(z), let f = PE0. Then

The only holomorphic function satisfying this condition is the constant function. So

with λ = 0 or 1. Since E0 is cyclic, it follows that P = 0 or I.

By the Stone–von Neumann theorem there is a unitary operator from L2(R) onto , unique up to multiplication by a scalar, intertwining the two representations of the Weyl commutation relations. By Schur's lemma and the Gelfand–Naimark construction, the matrix coefficient of any vector determines the vector up to a scalar multiple. Since the matrix coefficients of F = E0 and f = H0 are equal, it follows that the unitary is uniquely determined by the properties

and

Hence for f in L2(R)

so that

where

The operator is called the Segal–Bargmann transform[18] and B is called the Bargmann kernel.[19]

The adjoint of is given by the formula:

Fock model[edit]

The action of SU(1,1) on holomorphic Fock space was described by Bargmann (1970) and Itzykson (1967).

The metaplectic double cover of SU(1,1) can be constructed explicitly as pairs (g, γ) with

and

If g = g1g2, then

using the power series expansion of (1 + z)1/2 for |z| < 1.

The metaplectic representation is a unitary representation π(g, γ) of this group satisfying the covariance relations

where

Since is a reproducing kernel Hilbert space, any bounded operator T on it corresponds to a kernel given by a power series of its two arguments. In fact if

and F in , then

The covariance relations and analyticity of the kernel imply that for S = π(g, γ),

for some constant C. Direct calculation shows that

leads to an ordinary representation of the double cover.[20]

Coherent states can again be defined as the orbit of E0 under the metaplectic group.

For w complex, set

Then if and only if |w| < 1. In particular F0 = 1 = E0. Moreover,

where

Similarly the functions zFw lie in and form an orbit of the metaplectic group:

Since (Fw, E0) = 1, the matrix coefficient of the function E0 = 1 is given by[21]

Disk model[edit]

The projective representation of SL(2,R) on L2(R) or on break up as a direct sum of two irreducible representations, corresponding to even and odd functions of x or z. The two representations can be realized on Hilbert spaces of holomorphic functions on the unit disk; or, using the Cayley transform, on the upper half plane.[22][23]

The even functions correspond to holomorphic functions F+ for which

is finite; and the odd functions to holomorphic functions F for which

is finite. The polarized forms of these expressions define the inner products.

The action of the metaplectic group is given by

Irreducibility of these representations is established in a standard way.[24] Each representation breaks up as a direct sum of one dimensional eigenspaces of the rotation group each of which is generated by a C vector for the whole group. It follows that any closed invariant subspace is generated by the algebraic direct sum of eigenspaces it contains and that this sum is invariant under the infinitesimal action of the Lie algebra . On the other hand, that action is irreducible.

The isomorphism with even and odd functions in can be proved using the Gelfand–Naimark construction since the matrix coefficients associated to 1 and z in the corresponding representations are proportional. Itzykson (1967) gave another method starting from the maps

from the even and odd parts to functions on the unit disk. These maps intertwine the actions of the metaplectic group given above and send zn to a multiple of wn. Stipulating that U± should be unitary determines the inner products on functions on the disk, which can expressed in the form above.[25]

Although in these representations the operator L0 has positive spectrum—the feature that distinguishes the holomorphic discrete series representations of SU(1,1)—the representations do not lie in the discrete series of the metaplectic group. Indeed, Kashiwara & Vergne (1978) noted that the matrix coefficients are not square integrable, although their third power is.[26]

Harmonic oscillator and Hermite functions[edit]

Consider the following subspace of L2(R):

The operators

act on X is called the annihilation operator and Y the creation operator. They satisfy

Define the functions

We claim they are the eigenfunctions of the harmonic oscillator, D. To prove this we use the commutation relations above:

Next we have:

This is known for n = 0 and the commutation relation above yields

The nth Hermite function is defined by

pn is called the nth Hermite polynomial.

Let

Thus

The operators P, Q or equivalently A, A* act irreducibly on by a standard argument.[27][28]

Indeed, under the unitary isomorphism with holomorphic Fock space can be identified with C[z], the space of polynomials in z, with

If a subspace invariant under A and A* contains a non-zero polynomial p(z), then, applying a power of A*, it contains a non-zero constant; applying then a power of A, it contains all zn.

Under the isomorphism Fn is sent to a multiple of zn and the operator D is given by

Let

so that

In the terminology of physics A, A* give a single boson and L0 is the energy operator. It is diagonalizable with eigenvalues 1/2, 1, 3/2, ...., each of multiplicity one. Such a representation is called a positive energy representation.

Moreover,

so that the Lie bracket with L0 defines a derivation of the Lie algebra spanned by A, A* and I. Adjoining L0 gives the semidirect product. The infinitesimal version of the Stone–von Neumann theorem states that the above representation on C[z] is the unique irreducible positive energy representation of this Lie algebra with L0 = A*A + 1/2. For A lowers energy and A* raises energy. So any lowest energy vector v is annihilated by A and the module is exhausted by the powers of A* applied to v. It is thus a non-zero quotient of C[z] and hence can be identified with it by irreducibility.

Let

so that

These operators satisfy:

and act by derivations on the Lie algebra spanned by A, A* and I.

They are the infinitesimal operators corresponding to the metaplectic representation of SU(1,1).

The functions Fn are defined by

It follows that the Hermite functions are the orthonormal basis obtained by applying the Gram-Schmidt orthonormalization process to the basis xn exp -x2/2 of .

The completeness of the Hermite functions follows from the fact that the Bargmann transform is unitary and carries the orthonormal basis en(z) of holomorphic Fock space onto the Hn(x).

The heat operator for the harmonic oscillator is the operator on L2(R) defined as the diagonal operator

It corresponds to the heat kernel given by Mehler's formula:

This follows from the formula

To prove this formula note that if s = σ2, then by Taylor's formula

Thus Fσ,x lies in holomorphic Fock space and

an inner product that can be computed directly.

Wiener (1933, pp. 51–67) establishes Mehler's formula directly and uses a classical argument to prove that

tends to f in L2(R) as t decreases to 0. This shows the completeness of the Hermite functions and also, since

can be used to derive the properties of the Fourier transform.

There are other elementary methods for proving the completeness of the Hermite functions, for example using Fourier series.[29]

Sobolev spaces[edit]

The Sobolev spaces Hs, sometimes called Hermite-Sobolev spaces, are defined to be the completions of with respect to the norms

where

is the expansion of f in Hermite functions.[30]

Thus

The Sobolev spaces are Hilbert spaces. Moreover, Hs and Hs are in duality under the pairing

For s ≥ 0,

for some positive constant Cs.

Indeed, such an inequality can be checked for creation and annihilation operators acting on Hermite functions Hn and this implies the general inequality.[31]

It follows for arbitrary s by duality.

Consequently, for a quadratic polynomial R in P and Q

The Sobolev inequality holds for f in Hs with s > 1/2:

for any k ≥ 0.

Indeed, the result for general k follows from the case k = 0 applied to Qkf.

For k = 0 the Fourier inversion formula

implies

If s < t, the diagonal form of D, shows that the inclusion of Ht in Hs is compact (Rellich's lemma).

It follows from Sobolev's inequality that the intersection of the spaces Hs is . Functions in are characterized by the rapid decay of their Hermite coefficients an.

Standard arguments show that each Sobolev space is invariant under the operators W(z) and the metaplectic group.[32] Indeed, it is enough to check invariance when g is sufficiently close to the identity. In that case

with D + A an isomorphism from to

It follows that

If then

where the derivatives lie in

Similarly the partial derivatives of total degree k of U(s)V(t)f lie in Sobolev spaces of order sk/2.

Consequently, a monomial in P and Q of order 2k applied to f lies in Hsk and can be expressed as a linear combination of partial derivatives of U(s)V(t)f of degree ≤ 2k evaluated at 0.

Smooth vectors[edit]

The smooth vectors for the Weyl commutation relations are those u in L2(R) such that the map

is smooth. By the uniform boundedness theorem, this is equivalent to the requirement that each matrix coefficient (W(z)u,v) be smooth.

A vector is smooth if and only it lies in .[33] Sufficiency is clear. For necessity, smoothness implies that the partial derivatives of W(z)u lie in L2(R) and hence also Dku for all positive k. Hence u lies in the intersection of the Hk, so in .

It follows that smooth vectors are also smooth for the metaplectic group.

Moreover, a vector is in if and only if it is a smooth vector for the rotation subgroup of SU(1,1).

Analytic vectors[edit]

If Π(t) is a one parameter unitary group and for f in

then the vectors Π(f)ξ form a dense set of smooth vectors for Π.

In fact taking

the vectors v = Π(fε)ξ converge to ξ as ε decreases to 0 and

is an analytic function of t that extends to an entire function on C.

The vector is called an entire vector for Π.

The wave operator associated to the harmonic oscillator is defined by

The operator is diagonal with the Hermite functions Hn as eigenfunctions:

Since it commutes with D, it preserves the Sobolev spaces.

The analytic vectors constructed above can be rewritten in terms of the Hermite semigroup as

The fact that v is an entire vector for Π is equivalent to the summability condition

for all r > 0.

Any such vector is also an entire vector for U(s)V(t), that is the map

defined on R2 extends to an analytic map on C2.

This reduces to the power series estimate

So these form a dense set of entire vectors for U(s)V(t); this can also be checked directly using Mehler's formula.

The spaces of smooth and entire vectors for U(s)V(t) are each by definition invariant under the action of the metaplectic group as well as the Hermite semigroup.

Let

be the analytic continuation of the operators W(x,y) from R2 to C2 such that

Then W leaves the space of entire vectors invariant and satisfies

Moreover, for g in SL(2,R)

using the natural action of SL(2,R) on C2.

Formally

Oscillator semigroup[edit]

There is a natural double cover of the Olshanski semigroup H, and its closure that extends the double cover of SU(1,1) corresponding to the metaplectic group. It is given by pairs (g, γ) where g is an element of H or its closure

and γ is a square root of a.

Such a choice determines a unique branch of

for |z| < 1.

The unitary operators π(g) for g in SL(2,R) satisfy