Fact-checked by Grok 2 weeks ago

Projective representation

In the field of , a projective representation of a group G on a finite-dimensional V over a K is defined as a P: G \to \mathrm{PGL}(V), where \mathrm{PGL}(V) denotes the , obtained as the quotient of the \mathrm{GL}(V) by the of scalar multiples of the . Equivalently, it consists of a map \rho: G \to \mathrm{GL}(V) together with a function \alpha: G \times G \to K^\times (a 2-cocycle) satisfying \rho(g) \rho(h) = \alpha(g,h) \rho(gh) for all g, h \in G, and \rho(e) = I for the e \in G. Projective representations generalize ordinary linear representations, in which the cocycle \alpha is trivial (i.e., \alpha(g,h) = 1 for all g,h), and play a fundamental role in algebraic structures such as central extensions and twisted group algebras K_\alpha G. The cohomology class of \alpha lies in the second group H^2(G, K^\times), which measures the extent to which a projective fails to lift to a linear one; for K = \mathbb{C}, this is the of G. Every projective representation lifts to an ordinary linear representation of a representation group—a finite central extension \tilde{G} of G with isomorphic to the —allowing the study of projective representations through linear ones. For example, the of the S_n (with n \geq 4) is \mathbb{Z}/2\mathbb{Z}, yielding two non-isomorphic representation groups except for n=6. The concept was introduced by in 1904 as part of his foundational work on the representations of finite groups, including studies of symmetric and alternating groups, and further developed in his papers through 1911. Key results include Clifford's theorem (1937), which extends irreducible linear representations of subgroups to projective ones, and Mackey's refinement (1958) for general projective cases. These ideas have applications in the , where Schur multipliers inform the structure of groups. In , projective representations are crucial because physical states are elements of the (rays modulo phase factors), so symmetry groups act via projective unitary representations rather than ordinary ones. A prominent example is the rotation group \mathrm{SO}(3), whose projective representations correspond to spin systems, such as particles transforming under the double cover \mathrm{SU}(2); these yield dimensions $2j+1 for j. Bargmann's theorem (1954) establishes that continuous Lie groups admitting projective unitary representations possess central extensions by \mathrm{U}(1) to which these lift as ordinary unitary representations, facilitating the classification of particle representations under the .

Definitions and Foundations

Linear and Projective Representations

A linear representation of a group G over a field K is a homomorphism \rho: G \to \mathrm{GL}(V), where V is a vector space over K and \mathrm{GL}(V) is the general linear group consisting of all invertible linear transformations of V. This homomorphism preserves the group operation exactly, satisfying \rho(gh) = \rho(g)\rho(h) for all g, h \in G. The dimension of V is called the degree of the representation. In contrast, a projective representation of G on V is a P: G \to \mathrm{PGL}(V), where \mathrm{PGL}(V) = \mathrm{GL}(V)/K^\times is the , obtained by quotienting \mathrm{GL}(V) by the scalar multiples of the identity (with K^\times the of nonzero elements of K). Equivalently, it can be described via operators in \mathrm{GL}(V) satisfying P(g)P(h) = c(g,h) P(gh) for some scalar c(g,h) \in K^\times, where the projective equivalence ignores these phase factors. Thus, projective representations arise naturally as linear representations modulo scalars, capturing actions up to overall scaling. The study of projective representations originated in the early 20th century with Issai Schur's work on the irreducible representations of symmetric and alternating groups, where projective aspects emerged as essential for completeness during the period 1904–1911. Schur's investigations, including the construction of representation groups for these permutation groups, highlighted how projective representations extend the linear framework to address certain irreducibility phenomena. Projective representations of G can often be lifted to linear representations of a central extension of G, thereby reducing the analysis to standard linear cases. The obstruction to such a direct lift from projective to linear on G itself is measured by elements of the second cohomology group H^2(G, K^\times).

Description via Cocycles and

A projective representation \rho: G \to \mathrm{GL}(V) of a group G on a V over a F can be formalized using a 2-cocycle c: G \times G \to F^\times, where F^\times denotes the of the field. The function c satisfies the 2-cocycle condition c(g,h) \, c(gh,k) = c(g,hk) \, c(h,k) for all g, h, k \in G, and the representation obeys \rho(g) \rho(h) = c(g,h) \, \rho(gh) for all g, h \in G. This scalar factor c(g,h) accounts for the ambiguity inherent in projective representations. In the framework of group cohomology, with F^\times viewed as a trivial G-module, the equivalence classes of projective representations correspond bijectively to elements of the second cohomology group H^2(G, F^\times) = Z^2(G, F^\times) / B^2(G, F^\times), where Z^2(G, F^\times) is the group of 2-cocycles and B^2(G, F^\times) is the subgroup of 2-coboundaries. The Schur multiplier of G is precisely this group H^2(G, F^\times), which classifies the possible scalar factors up to equivalence. For example, over \mathbb{C}, the Schur multiplier of the symmetric group S_n for n \geq 4 is \mathbb{Z}/2\mathbb{Z}, yielding one nontrivial cohomology class, while for the alternating group A_n, it is \mathbb{Z}/2\mathbb{Z} for n=4,5 and n \geq 8, and \mathbb{Z}/6\mathbb{Z} for n=6,7. Two 2-cocycles c, c': G \times G \to F^\times define equivalent projective representations if they are cohomologous, meaning there exists a 1-cochain b: G \to F^\times such that c'(g,h) = \frac{b(g) \, b(h)}{b(gh)} \, c(g,h) for all g, h \in G; here, \frac{b(g) \, b(h)}{b(gh)} is the coboundary \delta b. This equivalence relation identifies cocycles within the same class \in H^2(G, F^\times). The cohomology class $$ provides an obstruction to lifting the projective representation to an ordinary linear representation of G: such a lift exists if and only if = 0 in H^2(G, F^\times), in which case the cocycle is a coboundary and the scalars can be absorbed into a genuine representation. Linear representations arise as the special case where c(g,h) = 1 for all g,h \in G.

Examples for Finite Groups

Representations of Finite Abelian Groups

For finite abelian groups G, the classification of projective representations over \mathbb{C} leverages the explicit structure of the second group H^2(G, \mathbb{C}^*). This group is isomorphic to \operatorname{Hom}(\wedge^2 G, \mathbb{C}^*), where \wedge^2 G denotes the exterior square of G. The isomorphism arises because the M(G) = H_2(G, \mathbb{Z}) \cong \wedge^2 G for abelian G, and by the universal coefficient theorem, H^2(G, \mathbb{C}^*) \cong \operatorname{Hom}(H_2(G, \mathbb{Z}), \mathbb{C}^*) since the \operatorname{Ext} term vanishes for the divisible module \mathbb{C}^*. This identification allows cocycles to be described as alternating bilinear forms on G with values in \mathbb{C}^*, facilitating explicit computations of projective structures. A simple illustration occurs for G = \mathbb{Z}/2\mathbb{Z}. Here, \wedge^2 G = 0 since G is cyclic of prime order, yielding H^2(G, \mathbb{C}^*) = 1, the . Consequently, every projective of G is equivalent to an ordinary linear . The irreducible projective coincide with the one-dimensional linear characters: the trivial , where the acts as the , and the sign , where it acts by by -1. For a case exhibiting nontrivial projective structure, consider G = \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}, the . The exterior square \wedge^2 G \cong \mathbb{Z}/2\mathbb{Z}, so H^2(G, \mathbb{C}^*) \cong \mathbb{Z}/2\mathbb{Z}. The nontrivial class corresponds to central extensions of G by \mathbb{Z}/2\mathbb{Z}, such as the D_4 of order 8. This group admits a faithful irreducible representation on \mathbb{C}^2 given by matrices \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}, where the generators of G map to \pm \sigma_x, \pm \sigma_z (up to the central scalars \pm 1), and the product generator maps to \pm i \sigma_y with \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}. This yields a two-dimensional irreducible projective representation of G, where the cocycle takes values in \{ \pm 1 \} to account for the phase factors. In contrast to linear representations of abelian groups, which are completely reducible into one-dimensional irreducibles, such projective irreducibles can have higher dimension determined by the order of the extension kernel. More generally, for a finite abelian G, while all irreducible linear representations over \mathbb{C} are one-dimensional (corresponding to characters of G), irreducible projective representations associated to nontrivial cocycles in H^2(G, \mathbb{C}^*) can have dimensions greater than one. These dimensions divide |G| and reflect the structure of central extensions classified by the ; for elementary abelian p-groups of r, the maximal dimension is p^{\lfloor r/2 \rfloor} for odd p. Higher-dimensional projective representations thus arise naturally from nontrivial elements of \operatorname{Hom}(\wedge^2 G, \mathbb{C}^*), enabling constructions beyond the linear case. Projective representations of finite abelian groups connect to an extended , where a projective character is the \chi: G \to \mathbb{C} of a projective representation. Unlike ordinary characters, these satisfy modified relations incorporating the cocycle, such as \sum_{g \in G} \overline{\omega(g)} \chi_1(g) \chi_2(g^{-1}) = |G| \delta_{\rho_1, \rho_2} for projective representations \rho_1, \rho_2 with multiplier \omega \in Z^2(G, \mathbb{C}^*). This framework generalizes classical to accommodate the cohomological data.

Discrete Fourier Transform Example

A concrete example of a projective representation arises for the finite G = \mathbb{Z}/p\mathbb{Z} \times \mathbb{Z}/p\mathbb{Z}, where p is prime, realized on the L^2(\mathbb{Z}/p\mathbb{Z}) of dimension p using operators inspired by the . Identify elements of G as pairs (a, b) with a, b \in \mathbb{Z}/p\mathbb{Z}. The translation operator T_a acts on the standard \{ |x\rangle \mid x \in \mathbb{Z}/p\mathbb{Z} \} by T_a |x\rangle = |x + a\rangle, implementing shifts in the "position" basis. The boost operator S_b acts by multiplication in the Fourier domain, S_b |x\rangle = \omega^{b x} |x\rangle, where \omega = e^{2\pi i / p} is a primitive p-th . These operators satisfy the projective relation T_a S_b = \omega^{-a b} S_b T_a, defining a projective unitary representation \rho: G \to \mathrm{PU}(p) via \rho(a, b) = T_a S_b. The phase factor \omega^{-a b} corresponds to the 2-cocycle c((a, 0), (0, b)) = \omega^{-a b} (extended bilinearly to general elements), which is nontrivial as a p-th . This representation is irreducible, as the operators generate the full algebra M_p(\mathbb{C}) through their Weyl-Heisenberg structure. The projective lifts to an ordinary linear of the central extension known as the over \mathbb{Z}/p\mathbb{Z}, a of order p^3 and 3 (as a over \mathbb{F}_p) with multiplication ((a, b, z), (a', b', z')) = (a + a', b + b', z + z' + a b'). The map sends (a, b) \mapsto T_a S_b e^{i \theta z} for central elements, preserving the group exactly. The \in H^2(G, \mathbb{U}(1)) \cong \mathbb{Z}/p\mathbb{Z} generates the group, confirming the cocycle is essentially nontrivial and classifies this extension up to equivalence. This example illustrates how discrete Fourier methods yield faithful projective structures for finite abelian groups beyond one .

Projective Representations of Lie Groups

Projective Representations of SO(3) and Spin

The second cohomology group H^2(\mathrm{SO}(3), \mathrm{U}(1)) is isomorphic to \mathbb{Z}/2\mathbb{Z}, indicating that projective representations of the rotation group SO(3) are classified into two equivalence classes: the ordinary (single-valued) representations and the projective (double-valued) ones corresponding to half-integer spin. The universal covering group of SO(3) is the special unitary group SU(2), which provides a 2:1 homomorphism \mathrm{SU}(2) \to \mathrm{SO}(3) that is a central extension by \mathbb{Z}/2\mathbb{Z}. The irreducible linear representations of SU(2), labeled by spin quantum numbers j = 0, 1/2, 1, 3/2, \dots, descend to representations of SO(3) via this covering map; those with integer j yield ordinary irreducible representations of SO(3), while those with half-integer j yield irreducible projective representations, differing by a sign under 360-degree rotations. A canonical example is the spin-1/2 projective representation on \mathbb{C}^2, realized using the \sigma_x, \sigma_y, \sigma_z. For a R by \theta around unit vector \mathbf{n}, the representation is given by \rho(R) = e^{-i (\theta/2) \mathbf{n} \cdot \boldsymbol{\sigma}}, where \boldsymbol{\sigma} = (\sigma_x, \sigma_y, \sigma_z). The composition of two such rotations satisfies \rho(R) \rho(S) = c(R,S) \rho(RS) with cocycle c(R,S) = \pm 1, reflecting the double-valued nature. In , representations describe fermions, such as electrons, where the projective nature enforces antisymmetric functions under 360-degree rotations, distinguishing them from bosons, which transform under integer- ordinary representations of SO(3). By Bargmann's theorem, all finite-dimensional unitary projective representations of SO(3) are equivalent to those obtained by projecting the finite-dimensional irreducible unitary representations of SU(2).

Universal Covers and Central Extensions

For a connected Lie group G, its universal covering group \hat{G} is a central extension of G by the \pi_1(G), with the covering map \hat{G} \to G having a discrete central isomorphic to \pi_1(G). Linear representations of \hat{G} descend to projective representations of G provided that the acts by scalar multiples on the representation , reflecting the phase factors inherent in projective actions. This systematically generates projective representations from ordinary ones on the cover, as the deck transformations introduce the necessary multipliers. More generally, projective representations of G can be realized as linear representations of a central extension H of G by the circle group U(1). Such an extension is constructed as H = \hat{G} \times_c U(1), where c: \hat{G} \times \hat{G} \to U(1) is a 2-cocycle representing a class in the second group H^2(\hat{G}, U(1)), and the group law on H incorporates the cocycle via (g_1, z_1)(g_2, z_2) = (g_1 g_2, z_1 z_2 c(g_1, g_2)). The cohomology class $$ classifies the possible extensions, and unitary representations of H project to projective unitary representations of G. A concrete instance of this arises with the double cover SU(2) \to SO(3), where linear spin representations on the cover yield projective representations on the . An illustrative example beyond the rotation group occurs with G = \mathrm{SL}(2, \mathbb{R}), whose universal cover \widetilde{\mathrm{SL}}(2, \mathbb{R}) is infinite-sheeted due to \pi_1(\mathrm{SL}(2, \mathbb{R})) \cong \mathbb{Z}. Linear representations of this cover descend to projective representations of \mathrm{SL}(2, \mathbb{R}), which is the double cover of the \mathrm{SO}(2,1) and thus relevant for transformations in two-dimensional . In general, a projective unitary representation of G lifts to a linear unitary representation of the central extension if the associated cocycle is cohomologous to a coboundary in the cover, meaning it arises from a 1-cochain on \hat{G}. For simply connected \hat{G}, the second H^2(\mathfrak{g}, \mathbb{R}) is often trivial—particularly for semisimple algebras—facilitating such lifts by ensuring that smooth central extensions are topologically trivial.

Finite-Dimensional Unitary Projective Representations

A finite-dimensional unitary projective representation of a G is a map \rho: G \to \mathrm{U}(H), where H is a finite-dimensional complex and \mathrm{U}(H) is the on H, such that \rho(g)^*\rho(g) = I for all g \in G (unitarity up to phase is automatic since phases are in \mathrm{U}(1)), and \rho(gh) = c(g,h) \rho(g) \rho(h) for all g, h \in G, where c: G \times G \to \mathrm{U}(1) is a measurable cocycle satisfying the 2-cocycle condition. Such representations arise naturally in , where observables transform linearly but states transform projectively under symmetries. For an irreducible finite-dimensional unitary projective representation over \mathbb{C}, extends to assert that the algebra of intertwiners consists precisely of scalar multiples by elements of \mathrm{U}(1). Moreover, one can select a equivalent representative \tilde{\rho} such that \det \tilde{\rho}(g) = 1 for all g \in G, meaning the image lies in the \mathrm{SL}(H); this is achieved by adjusting the cocycle by a suitable 1-coboundary to compensate for the phase, which lies in \mathrm{U}(1) due to unitarity. This normalization simplifies analysis and ensures the representation factors through a projective . In finite dimensions, every unitary projective representation of G is equivalent to a genuine linear unitary representation of a central extension \tilde{G} of G by \mathrm{U}(1), obtained via de-projectivization: specifically, the cocycle c defines a central extension $1 \to \mathrm{U}(1) \to \tilde{G} \to G \to 1, and there exists a unitary representation \tilde{\rho}: \tilde{G} \to \mathrm{U}(H) lifting \rho. For a concrete gauge fixing, one may choose representatives such that \operatorname{Tr}(\rho(g)) is real and fixed (e.g., via a "trace-zero" condition adjusted for the identity), ensuring uniqueness up to equivalence. This process is always possible in finite dimensions, contrasting with infinite-dimensional cases where obstructions may arise. For compact Lie groups, the classification of finite-dimensional irreducible unitary projective representations follows from the Peter-Weyl theorem and covering group theory: all such irreducibles arise as linear irreducible unitary representations of the universal cover \tilde{G} of G, restricted to those that are non-trivial on the kernel of the covering map (or tensor products with one-dimensional representations of if needed to match the cocycle class). The irreducibles of \tilde{G} are finite-dimensional and completely classified by highest weights, and projecting down yields the projective irreducibles of G. A canonical example is the rotation group \mathrm{SO}(3), whose universal cover is \mathrm{SU}(2); the finite-dimensional irreducible representations of \mathrm{SU}(2) (labeled by s j = 0, 1/2, 1, \dots, of dimension $2j+1) project to linear representations of \mathrm{SO}(3) precisely when j is integer, while half-integer j yield irreducible projective representations of \mathrm{SO}(3), essential for describing in .

Infinite-Dimensional Unitary Projective Representations: The Heisenberg Case

The H_3(\mathbb{R}) is defined as the group of $3 \times 3 upper triangular matrices with ones on the diagonal, parametrized by elements (x, y, z) \in \mathbb{R}^3 under the multiplication rule (x_1, y_1, z_1) \cdot (x_2, y_2, z_2) = (x_1 + x_2, y_1 + y_2, z_1 + z_2 + x_1 y_2). This structure makes H_3(\mathbb{R}) a central extension of the \mathbb{R}^2 (identified with the by the center Z = \{ (0,0,z) \} \cong \mathbb{R}) by \mathbb{R}, where the [(x,y,z), (x',y',z')] = (0,0, x y' - y x') lies in the center. A key example of an infinite-dimensional unitary projective representation arises from the Schrödinger representation of H_3(\mathbb{R}) on the L^2(\mathbb{R}, dx). This representation is realized via unitary operators consisting of translations T_a f(x) = f(x - a) for a \in \mathbb{R} and modulations M_b f(x) = e^{i b x} f(x) for b \in \mathbb{R}, extended to the full by \pi(x,y,z) f(t) = e^{i (z + x y / 2)} M_y T_x f(t). Projecting to the \mathbb{R}^2, these induce a projective representation where the operators satisfy the Weyl relation T_a M_b = e^{i a b} M_b T_a, corresponding to a nontrivial 2-cocycle \omega((a,0),(0,b)) = e^{i a b} on \mathbb{R}^2 with values in U(1). The asserts that every infinite-dimensional irreducible unitary of H_3(\mathbb{R}) with nontrivial central (i.e., where the center acts by a non-identity ) is unitarily equivalent to this Schrödinger representation. Unitarity is preserved on L^2(\mathbb{R}, dx) with respect to the , as both T_a and M_b are isometric isomorphisms. However, the induced cocycle on the quotient \mathbb{R}^2 is cohomologically nontrivial, lying in H^2(\mathbb{R}^2, U(1)) \cong \mathbb{R}, which prevents it from lifting to a true of \mathbb{R}^2 without the central extension. This construction provides the foundational infinite-dimensional unitary projective representation for the and serves as a basis for formulations in . It extends discrete analogs, such as the finite , to the continuous setting.

Infinite-Dimensional Unitary Projective Representations: Bargmann's Theorem

Bargmann's theorem provides a fundamental criterion for lifting projective unitary representations of connected s to linear unitary representations, particularly in infinite-dimensional s relevant to . Specifically, for a connected G with \mathfrak{g}, if the second continuous group H^2(\mathfrak{g}, \mathbb{R}) = 0, then every strongly continuous projective unitary of G on a separable lifts to a linear unitary of the universal \hat{G}. This result ensures that under the vanishing condition, the phase factors defining the projective can be absorbed into a true of the covering group, avoiding the need for additional central extensions. The proof relies on the and properties of the representations. A projective is described by a continuous 2-cocycle with values in U(1), which corresponds to a cocycle via the . Since H^2(\mathfrak{g}, \mathbb{R}) = 0, this cocycle is a coboundary, allowing the construction of a linear representation on the level. Strongly continuous extension to the group level then follows using the universal cover \hat{G}, where the cocycle trivializes globally due to the simply connected nature of \hat{G}. This approach highlights the role of continuous in de-projectivizing representations without introducing non-trivial phases. When H^2(\mathfrak{g}, \mathbb{R}) \neq 0, such as for certain Lie algebras underlying groups like the , lifts generally require non-trivial central extensions, potentially infinite-dimensional in the infinite-dimensional setting, and cannot be achieved solely via the universal cover. For instance, the Heisenberg case illustrates this non-vanishing , necessitating a central extension to realize projective representations linearly. The theorem extends naturally to infinite-dimensional contexts for strongly continuous representations, confirming that quantum symmetries of simply connected Lie groups with vanishing H^2 admit linear lifts on Hilbert spaces, a key insight for applications in . Historically, Bargmann's 1954 work built on Wigner's earlier investigations into ray representations in , providing a rigorous framework for when projective symmetries can be linearized, with later extensions to Banach spaces and infinite-dimensional groups appearing in post-2000 literature.

References

  1. [1]
    [PDF] Projective representations of groups - Alistair Savage
    We present an introduction to the basic concepts of projective representations of groups and representation groups, and discuss their relations with group ...
  2. [2]
    [PDF] Review of Projective Representations of Finite Groups
    Projective representations take their name from projective geometry. To be specific, let G be a finite group, K a field, and V a finite-dimensional.
  3. [3]
    Projective and spin representations - University of Alberta
    One motivation for the definition of projective representations comes from quantum mechanics. In quantum mechanics, the state of a system is specified by a ...
  4. [4]
    [PDF] Linear Representations of Groups - CIMAT
    Definition. A linear representation of the group G over the field K is. a homomorphism of G into the group GL(V ) of all invertible linear trans- formations ( ...
  5. [5]
    [PDF] Projective representations of groups - Alistair Savage
    Such a central extension reduces the problem of determining all projective representations of G to the determination of all linear representations of ˜G.
  6. [6]
    [PDF] An overview of the history of projective representations (spin ... - arXiv
    Dec 21, 2019 · ... symmetric groups and alternating groups ... [H2]. —, Work of Schur on the theory of group representations (trilogy of projective represen-.
  7. [7]
    [PDF] Group Cohomology, Projective Representations - Adam Monteleone
    Jul 22, 2022 · G to the projective linear group3 PGL(V ) = GL(V )/F∗ and V is a K−vector space. ... This gives d ways of defining a projective representation ...
  8. [8]
    [PDF] Schur Multipliers and Second Quandle Homology - arXiv
    Dec 11, 2018 · ... Schur multiplier1 of the underlying abelian group, that is, the exterior square of the group [Mil, NP]. In this paper, we demonstrate a new ...
  9. [9]
    [PDF] ON IRREDUCIBLE PROJECTIVE REPRESENTATIONS OF FINITE ...
    Abstract. The paper is a survey type article in which we present some results on irreducible projective representations of finite groups. Section 2 includes ...
  10. [10]
    Projective Representations of Finite Groups - Gregory Karpilovsky
    Offers an understanding of projective representations of finite groups for algebraists, number theorists, mathematical researchers studying modern algebra, and ...
  11. [11]
    [PDF] character theory for projective representations of finite groups
    Abstract. In this paper, we construct a character theory for projective representations of finite groups and deduce some consequences of this theory.
  12. [12]
    [PDF] arXiv:1701.07902v1 [quant-ph] 26 Jan 2017
    Jan 26, 2017 · A very important area of application for the Weyl–Heisenberg group is that of time-frequency analysis of signals; then the operators X and Z may ...
  13. [13]
    Some projective representations of finite abelian groups
    May 18, 2009 · Karpilovsky, G., Projective representations of finite groups(Marcel Dekker, 1985).Google Scholar. 3. 3. Morris, A. O., On a generalized ...
  14. [14]
  15. [15]
  16. [16]
  17. [17]
    Notes on $\widetilde{\mathrm{SL}}(2,\mathbb{R})$ representations
    ### Summary of Projective Representations from Universal Cover of SL(2,R)
  18. [18]
    [PDF] Projective unitary representations of infinite dimensional Lie groups
    Jan 19, 2015 · smooth projective representations. ... In Subsection 10.1, we consider abelian. Lie groups, whose central extensions are related to Heisenberg ...
  19. [19]
    None
    Summary of each segment:
  20. [20]
    [PDF] Projective Representations, Bogomolov Multiplier, and Their ... - arXiv
    Jul 16, 2025 · projective representation of SO(3), which is realized, for example ... A 2-cocycle α: G × G → U(1) satisfying the normalization ...
  21. [21]
    [PDF] Lectures and problems in representation theory - MIT Mathematics
    Apr 23, 2005 · standard 2-dimensional representation of SU(2). ... Since this action preserves the norm on V , we have a homomorphism h : SU(2) → SO(3).
  22. [22]