Projective representation

Last updated

In the field of representation theory in mathematics, a projective representation of a group G on a vector space V over a field F is a group homomorphism from G to the projective linear group

Contents

where GL(V) is the general linear group of invertible linear transformations of V over F, and F is the normal subgroup consisting of nonzero scalar multiples of the identity transformation (see Scalar transformation). [1]

In more concrete terms, a projective representation of is a collection of operators satisfying the homomorphism property up to a constant:

for some constant . Equivalently, a projective representation of is a collection of operators , such that . Note that, in this notation, is a set of linear operators related by multiplication with some nonzero scalar.

If it is possible to choose a particular representative in each family of operators in such a way that the homomorphism property is satisfied on the nose, rather than just up to a constant, then we say that can be "de-projectivized", or that can be "lifted to an ordinary representation". More concretely, we thus say that can be de-projectivized if there are for each such that . This possibility is discussed further below.

Linear representations and projective representations

One way in which a projective representation can arise is by taking a linear group representation of G on V and applying the quotient map

which is the quotient by the subgroup F of scalar transformations (diagonal matrices with all diagonal entries equal). The interest for algebra is in the process in the other direction: given a projective representation, try to 'lift' it to an ordinary linear representation. A general projective representation ρ: G → PGL(V) cannot be lifted to a linear representation G → GL(V), and the obstruction to this lifting can be understood via group cohomology, as described below.

However, one can lift a projective representation of G to a linear representation of a different group H, which will be a central extension of G. The group is the subgroup of defined as follows:

,

where is the quotient map of onto . Since is a homomorphism, it is easy to check that is, indeed, a subgroup of . If the original projective representation is faithful, then is isomorphic to the preimage in of .

We can define a homomorphism by setting . The kernel of is:

,

which is contained in the center of . It is clear also that is surjective, so that is a central extension of . We can also define an ordinary representation of by setting . The ordinary representation of is a lift of the projective representation of in the sense that:

.

If G is a perfect group there is a single universal perfect central extension of G that can be used.

Group cohomology

The analysis of the lifting question involves group cohomology. Indeed, if one fixes for each g in G a lifted element L(g) in lifting from PGL(V) back to GL(V), the lifts then satisfy

for some scalar c(g,h) in F. It follows that the 2-cocycle or Schur multiplier c satisfies the cocycle equation

for all g, h, k in G. This c depends on the choice of the lift L; a different choice of lift L(g) = f(g) L(g) will result in a different cocycle

cohomologous to c. Thus L defines a unique class in H2(G, F). This class might not be trivial. For example, in the case of the symmetric group and alternating group, Schur established that there is exactly one non-trivial class of Schur multiplier, and completely determined all the corresponding irreducible representations. [2]

In general, a nontrivial class leads to an extension problem for G. If G is correctly extended we obtain a linear representation of the extended group, which induces the original projective representation when pushed back down to G. The solution is always a central extension. From Schur's lemma, it follows that the irreducible representations of central extensions of G, and the irreducible projective representations of G, are essentially the same objects.

First example: discrete Fourier transform

Consider the field of integers mod , where is prime, and let be the -dimensional space of functions on with values in . For each in , define two operators, and on as follows:

We write the formula for as if and were integers, but it is easily seen that the result only depends on the value of and mod . The operator is a translation, while is a shift in frequency space (that is, it has the effect of translating the discrete Fourier transform of ).

One may easily verify that for any and in , the operators and commute up to multiplication by a constant:

.

We may therefore define a projective representation of as follows:

,

where denotes the image of an operator in the quotient group . Since and commute up to a constant, is easily seen to be a projective representation. On the other hand, since and do not actually commute—and no nonzero multiples of them will commute— cannot be lifted to an ordinary (linear) representation of .

Since the projective representation is faithful, the central extension of obtained by the construction in the previous section is just the preimage in of the image of . Explicitly, this means that is the group of all operators of the form

for . This group is a discrete version of the Heisenberg group and is isomorphic to the group of matrices of the form

with .

Projective representations of Lie groups

Studying projective representations of Lie groups leads one to consider true representations of their central extensions (see Group extension § Lie groups). In many cases of interest it suffices to consider representations of covering groups. Specifically, suppose is a connected cover of a connected Lie group , so that for a discrete central subgroup of . (Note that is a special sort of central extension of .) Suppose also that is an irreducible unitary representation of (possibly infinite dimensional). Then by Schur's lemma, the central subgroup will act by scalar multiples of the identity. Thus, at the projective level, will descend to . That is to say, for each , we can choose a preimage of in , and define a projective representation of by setting

,

where denotes the image in of an operator . Since is contained in the center of and the center of acts as scalars, the value of does not depend on the choice of .

The preceding construction is an important source of examples of projective representations. Bargmann's theorem (discussed below) gives a criterion under which every irreducible projective unitary representation of arises in this way.

Projective representations of SO(3)

A physically important example of the above construction comes from the case of the rotation group SO(3), whose universal cover is SU(2). According to the representation theory of SU(2), there is exactly one irreducible representation of SU(2) in each dimension. When the dimension is odd (the "integer spin" case), the representation descends to an ordinary representation of SO(3). [3] When the dimension is even (the "fractional spin" case), the representation does not descend to an ordinary representation of SO(3) but does (by the result discussed above) descend to a projective representation of SO(3). Such projective representations of SO(3) (the ones that do not come from ordinary representations) are referred to as "spinorial representations", whose elements (vectors) are called Spinors.

By an argument discussed below, every finite-dimensional, irreducible projective representation of SO(3) comes from a finite-dimensional, irreducible ordinary representation of SU(2).

Examples of covers, leading to projective representations

Notable cases of covering groups giving interesting projective representations:

Finite-dimensional projective unitary representations

In quantum physics, symmetry of a physical system is typically implemented by means of a projective unitary representation of a Lie group on the quantum Hilbert space, that is, a continuous homomorphism

where is the quotient of the unitary group by the operators of the form . The reason for taking the quotient is that physically, two vectors in the Hilbert space that are proportional represent the same physical state. [That is to say, the space of (pure) states is the set of equivalence classes of unit vectors, where two unit vectors are considered equivalent if they are proportional.] Thus, a unitary operator that is a multiple of the identity actually acts as the identity on the level of physical states.

A finite-dimensional projective representation of then gives rise to a projective unitary representation of the Lie algebra of . In the finite-dimensional case, it is always possible to "de-projectivize" the Lie-algebra representation simply by choosing a representative for each having trace zero. [4] In light of the homomorphisms theorem, it is then possible to de-projectivize itself, but at the expense of passing to the universal cover of . [5] That is to say, every finite-dimensional projective unitary representation of arises from an ordinary unitary representation of by the procedure mentioned at the beginning of this section.

Specifically, since the Lie-algebra representation was de-projectivized by choosing a trace-zero representative, every finite-dimensional projective unitary representation of arises from a determinant-one ordinary unitary representation of (i.e., one in which each element of acts as an operator with determinant one). If is semisimple, then every element of is a linear combination of commutators, in which case every representation of is by operators with trace zero. In the semisimple case, then, the associated linear representation of is unique.

Conversely, if is an irreducible unitary representation of the universal cover of , then by Schur's lemma, the center of acts as scalar multiples of the identity. Thus, at the projective level, descends to a projective representation of the original group . Thus, there is a natural one-to-one correspondence between the irreducible projective representations of and the irreducible, determinant-one ordinary representations of . (In the semisimple case, the qualifier "determinant-one" may be omitted, because in that case, every representation of is automatically determinant one.)

An important example is the case of SO(3), whose universal cover is SU(2). Now, the Lie algebra is semisimple. Furthermore, since SU(2) is a compact group, every finite-dimensional representation of it admits an inner product with respect to which the representation is unitary. [6] Thus, the irreducible projective representations of SO(3) are in one-to-one correspondence with the irreducible ordinary representations of SU(2).

Infinite-dimensional projective unitary representations: the Heisenberg case

The results of the previous subsection do not hold in the infinite-dimensional case, simply because the trace of is typically not well defined. Indeed, the result fails: Consider, for example, the translations in position space and in momentum space for a quantum particle moving in , acting on the Hilbert space . [7] These operators are defined as follows:

for all . These operators are simply continuous versions of the operators and described in the "First example" section above. As in that section, we can then define a projective unitary representation of :

because the operators commute up to a phase factor. But no choice of the phase factors will lead to an ordinary unitary representation, since translations in position do not commute with translations in momentum (and multiplying by a nonzero constant will not change this). These operators do, however, come from an ordinary unitary representation of the Heisenberg group, which is a one-dimensional central extension of . [8] (See also the Stone–von Neumann theorem.)

Infinite-dimensional projective unitary representations: Bargmann's theorem

On the other hand, Bargmann's theorem states that if the second Lie algebra cohomology group of is trivial, then every projective unitary representation of can be de-projectivized after passing to the universal cover. [9] [10] More precisely, suppose we begin with a projective unitary representation of a Lie group . Then the theorem states that can be lifted to an ordinary unitary representation of the universal cover of . This means that maps each element of the kernel of the covering map to a scalar multiple of the identity—so that at the projective level, descends to —and that the associated projective representation of is equal to .

The theorem does not apply to the group —as the previous example shows—because the second cohomology group of the associated commutative Lie algebra is nontrivial. Examples where the result does apply include semisimple groups (e.g., SL(2,R)) and the Poincaré group. This last result is important for Wigner's classification of the projective unitary representations of the Poincaré group.

The proof of Bargmann's theorem goes by considering a central extension of , constructed similarly to the section above on linear representations and projective representations, as a subgroup of the direct product group , where is the Hilbert space on which acts and is the group of unitary operators on . The group is defined as

As in the earlier section, the map given by is a surjective homomorphism whose kernel is so that is a central extension of . Again as in the earlier section, we can then define a linear representation of by setting . Then is a lift of in the sense that , where is the quotient map from to .

A key technical point is to show that is a Lie group. (This claim is not so obvious, because if is infinite dimensional, the group is an infinite-dimensional topological group.) Once this result is established, we see that is a one-dimensional Lie group central extension of , so that the Lie algebra of is also a one-dimensional central extension of (note here that the adjective "one-dimensional" does not refer to and , but rather to the kernel of the projection map from those objects onto and respectively). But the cohomology group may be identified with the space of one-dimensional (again, in the aforementioned sense) central extensions of ; if is trivial then every one-dimensional central extension of is trivial. In that case, is just the direct sum of with a copy of the real line. It follows that the universal cover of must be just a direct product of the universal cover of with a copy of the real line. We can then lift from to (by composing with the covering map) and finally restrict this lift to the universal cover of .

See also

Notes

  1. Gannon 2006 , pp. 176–179.
  2. Schur 1911
  3. Hall 2015 Section 4.7
  4. Hall 2013 Proposition 16.46
  5. Hall 2013 Theorem 16.47
  6. Hall 2015 proof of Theorem 4.28
  7. Hall 2013 Example 16.56
  8. Hall 2013 Exercise 6 in Chapter 14
  9. Bargmann 1954
  10. Simms 1971

Related Research Articles

<span class="mw-page-title-main">Lie algebra</span> Algebraic structure used in analysis

In mathematics, a Lie algebra is a vector space together with an operation called the Lie bracket, an alternating bilinear map , that satisfies the Jacobi identity. In other words, a Lie algebra is an algebra over a field for which the multiplication operation is alternating and satisfies the Jacobi identity. The Lie bracket of two vectors and is denoted . A Lie algebra is typically a non-associative algebra. However, every associative algebra gives rise to a Lie algebra, consisting of the same vector space with the commutator Lie bracket, .

<span class="mw-page-title-main">Quaternion group</span> Non-abelian group of order eight

In group theory, the quaternion group Q8 (sometimes just denoted by Q) is a non-abelian group of order eight, isomorphic to the eight-element subset of the quaternions under multiplication. It is given by the group presentation

<span class="mw-page-title-main">Representation of a Lie group</span> Group representation

In mathematics and theoretical physics, a representation of a Lie group is a linear action of a Lie group on a vector space. Equivalently, a representation is a smooth homomorphism of the group into the group of invertible operators on the vector space. Representations play an important role in the study of continuous symmetry. A great deal is known about such representations, a basic tool in their study being the use of the corresponding 'infinitesimal' representations of Lie algebras.

In mathematics, the Peter–Weyl theorem is a basic result in the theory of harmonic analysis, applying to topological groups that are compact, but are not necessarily abelian. It was initially proved by Hermann Weyl, with his student Fritz Peter, in the setting of a compact topological group G. The theorem is a collection of results generalizing the significant facts about the decomposition of the regular representation of any finite group, as discovered by Ferdinand Georg Frobenius and Issai Schur.

In mathematics, a Lie algebroid is a vector bundle together with a Lie bracket on its space of sections and a vector bundle morphism , satisfying a Leibniz rule. A Lie algebroid can thus be thought of as a "many-object generalisation" of a Lie algebra.

<span class="mw-page-title-main">Lie algebra representation</span>

In the mathematical field of representation theory, a Lie algebra representation or representation of a Lie algebra is a way of writing a Lie algebra as a set of matrices in such a way that the Lie bracket is given by the commutator. In the language of physics, one looks for a vector space together with a collection of operators on satisfying some fixed set of commutation relations, such as the relations satisfied by the angular momentum operators.

<span class="mw-page-title-main">Irreducible representation</span> Type of group and algebra representation

In mathematics, specifically in the representation theory of groups and algebras, an irreducible representation or irrep of an algebraic structure is a nonzero representation that has no proper nontrivial subrepresentation , with closed under the action of .

<span class="mw-page-title-main">Circle group</span> Lie group of complex numbers of unit modulus; topologically a circle

In mathematics, the circle group, denoted by or , is the multiplicative group of all complex numbers with absolute value 1, that is, the unit circle in the complex plane or simply the unit complex numbers

In mathematics, Schur's lemma is an elementary but extremely useful statement in representation theory of groups and algebras. In the group case it says that if M and N are two finite-dimensional irreducible representations of a group G and φ is a linear map from M to N that commutes with the action of the group, then either φ is invertible, or φ = 0. An important special case occurs when M = N, i.e. φ is a self-map; in particular, any element of the center of a group must act as a scalar operator on M. The lemma is named after Issai Schur who used it to prove the Schur orthogonality relations and develop the basics of the representation theory of finite groups. Schur's lemma admits generalisations to Lie groups and Lie algebras, the most common of which are due to Jacques Dixmier and Daniel Quillen.

<span class="mw-page-title-main">Compact group</span> Topological group with compact topology

In mathematics, a compact (topological) group is a topological group whose topology realizes it as a compact topological space. Compact groups are a natural generalization of finite groups with the discrete topology and have properties that carry over in significant fashion. Compact groups have a well-understood theory, in relation to group actions and representation theory.

The representation theory of groups is a part of mathematics which examines how groups act on given structures.

In group theory, restriction forms a representation of a subgroup using a known representation of the whole group. Restriction is a fundamental construction in representation theory of groups. Often the restricted representation is simpler to understand. Rules for decomposing the restriction of an irreducible representation into irreducible representations of the subgroup are called branching rules, and have important applications in physics. For example, in case of explicit symmetry breaking, the symmetry group of the problem is reduced from the whole group to one of its subgroups. In quantum mechanics, this reduction in symmetry appears as a splitting of degenerate energy levels into multiplets, as in the Stark or Zeeman effect.

In theoretical physics and mathematics, a Wess–Zumino–Witten (WZW) model, also called a Wess–Zumino–Novikov–Witten model, is a type of two-dimensional conformal field theory named after Julius Wess, Bruno Zumino, Sergei Novikov and Edward Witten. A WZW model is associated to a Lie group, and its symmetry algebra is the affine Lie algebra built from the corresponding Lie algebra. By extension, the name WZW model is sometimes used for any conformal field theory whose symmetry algebra is an affine Lie algebra.

<span class="mw-page-title-main">Representation theory of the Lorentz group</span> Representation of the symmetry group of spacetime in special relativity

The Lorentz group is a Lie group of symmetries of the spacetime of special relativity. This group can be realized as a collection of matrices, linear transformations, or unitary operators on some Hilbert space; it has a variety of representations. This group is significant because special relativity together with quantum mechanics are the two physical theories that are most thoroughly established, and the conjunction of these two theories is the study of the infinite-dimensional unitary representations of the Lorentz group. These have both historical importance in mainstream physics, as well as connections to more speculative present-day theories.

<span class="mw-page-title-main">Particle physics and representation theory</span> Physics-mathematics connection

There is a natural connection between particle physics and representation theory, as first noted in the 1930s by Eugene Wigner. It links the properties of elementary particles to the structure of Lie groups and Lie algebras. According to this connection, the different quantum states of an elementary particle give rise to an irreducible representation of the Poincaré group. Moreover, the properties of the various particles, including their spectra, can be related to representations of Lie algebras, corresponding to "approximate symmetries" of the universe.

In physics and particularly in particle physics, a multiplet is the state space for 'internal' degrees of freedom of a particle, that is, degrees of freedom associated to a particle itself, as opposed to 'external' degrees of freedom such as the particle's position in space. Examples of such degrees of freedom are the spin state of a particle in quantum mechanics, or the color, isospin and hypercharge state of particles in the Standard model of particle physics. Formally, we describe this state space by a vector space which carries the action of a group of continuous symmetries.

<span class="mw-page-title-main">Representation theory</span> Branch of mathematics that studies abstract algebraic structures

Representation theory is a branch of mathematics that studies abstract algebraic structures by representing their elements as linear transformations of vector spaces, and studies modules over these abstract algebraic structures. In essence, a representation makes an abstract algebraic object more concrete by describing its elements by matrices and their algebraic operations. The theory of matrices and linear operators is well-understood, so representations of more abstract objects in terms of familiar linear algebra objects helps glean properties and sometimes simplify calculations on more abstract theories.

In algebraic geometry and differential geometry, the nonabelian Hodge correspondence or Corlette–Simpson correspondence is a correspondence between Higgs bundles and representations of the fundamental group of a smooth, projective complex algebraic variety, or a compact Kähler manifold.

This is a glossary of representation theory in mathematics.

<span class="mw-page-title-main">Representations of classical Lie groups</span>

In mathematics, the finite-dimensional representations of the complex classical Lie groups , , , , , can be constructed using the general representation theory of semisimple Lie algebras. The groups , , are indeed simple Lie groups, and their finite-dimensional representations coincide with those of their maximal compact subgroups, respectively , , . In the classification of simple Lie algebras, the corresponding algebras are

References