Fact-checked by Grok 2 weeks ago

Antilinear map

In mathematics, particularly in the context of complex vector spaces, an antilinear map (also known as a conjugate-linear map) is a function f: V \to W between two complex vector spaces V and W that preserves addition but conjugates the scalars in the homogeneity condition, satisfying f(\alpha v + \beta w) = \bar{\alpha} f(v) + \bar{\beta} f(w) for all vectors v, w \in V and complex scalars \alpha, \beta \in \mathbb{C}, where \bar{\cdot} denotes the complex conjugate. Antilinear maps form a natural counterpart to s in complex linear algebra and can be viewed as linear maps when the domain is equipped with the conjugate \alpha \cdot v = \bar{\alpha} v. The space of bounded antilinear operators on a X is often denoted B_a(X), and these operators are closed under addition and (with conjugation on the scalars). Key properties include the fact that the of two antilinear maps is linear, while the of an antilinear map with a linear map is antilinear; eigenvalues of bounded antilinear operators on Hilbert spaces lie on circles centered at the origin in the . Antilinear maps play a significant role in , where they arise in the study of numerical ranges, polar decompositions, and complex symmetric operators; for instance, the of a linear can induce antilinear structures in certain contexts. In , they appear in the of group representations and in the study of real structures on complex representations. Additionally, in on Hilbert spaces, normal antilinear operators admit decompositions into one-dimensional eigenspaces or two-dimensional invariant subspaces, as per extensions of the . Their applications extend to theory for modeling operations like time reversal, though these are rooted in mathematical frameworks such as Tomita-Takesaki theory for algebras.

Definition and Basics

Formal Definition

In the context of complex vector spaces, an antilinear map (also known as a conjugate-linear map) is a T: V \to W between two vector spaces V and W over the field \mathbb{C} that preserves but conjugates . Specifically, it satisfies the T(\alpha v + \beta w) = \bar{\alpha} T(v) + \bar{\beta} T(w) for all scalars \alpha, \beta \in \mathbb{C} and all vectors v, w \in V, where \bar{\alpha} and \bar{\beta} denote the complex conjugates of \alpha and \beta, respectively. This definition presupposes that V and W are equipped with the standard structure of complex vector spaces, where involves complex numbers and the conjugation operation z \mapsto \bar{z} for z \in \mathbb{C} is the standard one satisfying \bar{z} = \operatorname{Re}(z) - i \operatorname{Im}(z). Antilinear maps differ from linear maps, which satisfy T(\alpha v + \beta w) = \alpha T(v) + \beta T(w) without conjugation. For contrast, sesquilinear forms extend this idea to bilinear-like structures that are linear in one argument and antilinear in the other.

Relation to Linear Maps

Antilinear maps differ from s primarily in their handling of over the numbers. A T: V \to W between complex vector spaces satisfies T(\alpha v) = \alpha T(v) for all scalars \alpha \in \mathbb{C} and vectors v \in V, preserving the action of complex scalars directly. In contrast, an antilinear map T satisfies T(\alpha v) = \bar{\alpha} T(v), where \bar{\alpha} denotes the of \alpha, effectively conjugating the scalar before applying the map. This distinction arises because antilinearity incorporates the of complex conjugation, which reverses the imaginary part of scalars. However, when restricted to real scalars \alpha \in \mathbb{R}, where \bar{\alpha} = \alpha, the behaviors of linear and antilinear maps coincide, making both additive and homogeneous over the reals. This connection to real linearity facilitates a deeper structural through the process of realification. For a complex vector space V, the underlying real vector space V_{\mathbb{R}} is obtained by restricting scalars to \mathbb{R}, effectively identifying \mathbb{C} with \mathbb{R}^2 via the basis \{1, i\}, doubling the over \mathbb{R}. An antilinear T: V \to W then induces a real-linear on V_{\mathbb{R}} \to W_{\mathbb{R}}, as it preserves addition and real while the conjugation on imaginary units aligns with the real structure (specifically, T(iv) = -i T(v), compatible with the real-linear action of multiplication by i as a real ). Equivalently, via the conjugate space V \mapsto \bar{V}, where vectors in \bar{V} carry conjugated scalars, an antilinear corresponds to a complex-linear \bar{V} \to W, providing an between the categories of antilinear and linear maps in this conjugated setting. This realification underscores that antilinearity is a natural extension of when viewing complex spaces through their real substructures. The concept of antilinear maps originated in the early , emerging in to preserve real-valued structures in holomorphic mappings and in physics to model symmetries like time reversal. In , Eugene P. Wigner introduced antilinear operators in to represent time-reversal invariance, where the antiunitary nature ensures reversal of momenta and spins while conjugating phases to maintain probability conservation. This foundational role in physics, alongside applications in spinor calculus and , highlighted antilinearity's utility in capturing operations that respect real geometries within complex frameworks.

Examples and Characterizations

Dual Space Applications

In the context of dual spaces for a complex vector space V, the anti-dual space V^{*\mathrm{anti}} consists of all antilinear functionals \phi: V \to \mathbb{C} satisfying \phi(\alpha v + \beta w) = \bar{\alpha} \phi(v) + \bar{\beta} \phi(w) for all \alpha, \beta \in \mathbb{C} and v, w \in V. This space V^{*\mathrm{anti}} forms a complex vector space under pointwise addition and scalar multiplication (\alpha \phi)(v) = \alpha \phi(v), with complex dimension \dim_{\mathbb{C}} V^{*\mathrm{anti}} = \dim_{\mathbb{C}} V. The motivation for considering such antilinear functionals arises prominently in physics, especially quantum mechanics, where sesquilinear inner products \langle \phi | \psi \rangle are linear in \psi and antilinear in \phi, allowing antilinear maps to represent dual elements like bras in the Dirac formalism or operators in time-reversal symmetry. As real vector spaces, V^{*\mathrm{anti}} is isomorphic to the dual of the underlying real vector space (V_{\mathbb{R}})^*, with both having real dimension $2 \dim_{\mathbb{C}} V. The explicit bijection identifies each antilinear \phi with the corresponding real-linear functional on V_{\mathbb{R}} via the decomposition \phi(v) = \mathrm{Re}(\phi(v)) + i \mathrm{Im}(\phi(v)), where \mathrm{Re}(\phi) and \mathrm{Im}(\phi) are \mathbb{R}-linear maps from V_{\mathbb{R}} to \mathbb{R} satisfying the compatibility conditions imposed by antilinearity, such as \phi(i v) = -i \phi(v). This isomorphism preserves the real dimension equality \dim_{\mathbb{R}} V^{*\mathrm{anti}} = \dim_{\mathbb{R}} V and facilitates viewing antilinear structures through real geometry. For a concrete example, consider V = \mathbb{C}^n with the standard basis \{e_1, \dots, e_n\}, where e_j has 1 in the j-th position and 0 elsewhere. The standard anti-linear dual basis \{\phi_1, \dots, \phi_n\} for V^{*\mathrm{anti}} is defined by \phi_j(e_k) = \delta_{jk} for j,k = 1, \dots, n, extending antilinearly to general vectors: for v = \sum_{k=1}^n z_k e_k with z_k \in \mathbb{C}, \phi_j(v) = \bar{z}_j. These basis functionals are linearly independent over \mathbb{C} and span V^{*\mathrm{anti}}, confirming the complex dimension n.

Physical Representations

In , the time-reversal T is a fundamental example of an antilinear map, acting on wave functions \psi in the by reversing momenta and incorporating complex conjugation to preserve the form of the under time reversal. Specifically, for a state \psi, the action satisfies T(i \psi) = -i T(\psi), reflecting its antilinearity due to the conjugation component, while maintaining anti-unitarity to ensure under the . This is typically expressed as T = U K, where K denotes complex conjugation in a chosen basis and U is a specifying the spatial reversal, such as U = i \sigma_y for particles. A concrete illustration occurs in the representation of particles, where the time-reversal operator squares to minus the : T^2 = -1. For a \psi = \begin{pmatrix} \alpha \\ \beta \end{pmatrix}, applying T yields T\psi = i \sigma_y \begin{pmatrix} \alpha^* \\ \beta^* \end{pmatrix} = \begin{pmatrix} i \beta^* \\ -i \alpha^* \end{pmatrix}, and iterating gives T^2 \psi = -\psi, which distinguishes fermionic systems and underlies phenomena like Kramers' degeneracy in time-reversal invariant systems without spin-orbit coupling. This property arises because the antilinear nature prevents T from being diagonalizable in the usual sense for spins, enforcing paired degenerate states. In , charge conjugation provides another physical realization of an antilinear map, particularly in its action on Dirac fields, where it interchanges particles and antiparticles while involving complex conjugation on the components. The on a \psi is given by \psi_c = C \overline{\psi}^T, with C = i \gamma^2 \gamma^0 in the Dirac representation (satisfying C^\dagger = C^{-1} = -C and C \gamma^\mu C^{-1} = -(\gamma^\mu)^T), explicitly incorporating \psi^* through the \overline{\psi} = \psi^\dagger \gamma^0. This antilinear form ensures the invariance of the under charge conjugation for massless fields or in contexts like Majorana fermions, swapping for particles and antiparticles in the quantized theory. For instance, in the quantized Dirac , the charge conjugation \hat{C} acts on the expansion \psi(x) = \int d^3 p \, [u(p) a_p e^{-ip \cdot x} + v(p) b_p^\dagger e^{ip \cdot x}] by \hat{C} \psi(x) \hat{C}^{-1} = \eta_C \overline{\psi}^T(x), where \eta_C is a , effectively exchanging a_p \leftrightarrow b_p up to signs and relying on the antilinear conjugation to map positive-energy solutions to negative-energy () ones. This structure is crucial for CPT invariance and for constructing charge-neutral states, such as in neutral pion decay processes. In , antilinear maps appear in the classification of irreducible representations of compact groups, such as SU(2), via the Frobenius-Schur indicator, which detects the existence of invariant sesquilinear forms compatible with an antilinear . For an \rho: G \to \mathrm{GL}(V) over \mathbb{C}, the indicator is defined as \nu(\rho) = \frac{1}{\dim V} \sum_{g \in G} \chi_\rho(g^2), where \chi_\rho is the ; values of +1, $0, or -1 indicate real, , or quaternionic types, respectively. The quaternionic case (\nu = -1) corresponds to the existence of a non-degenerate G-invariant antilinear map J: V \to V with J^2 = -1 and \rho(g) J = J \overline{\rho(g)} for all g \in G, intertwining the representation with its . For SU(2), the irreducible representations labeled by spin j (dimension $2j+1) have Frobenius-Schur indicator (-1)^{2j}: integer j yield +1 (real type, admitting a symmetric invariant ), while j yield -1 (quaternionic type, with an antisymmetric invariant preserved by an antilinear J). The fundamental representation (j = 1/2) exemplifies this, where J = i \sigma_y serves as the antilinear map, reflecting the pseudo-real nature of spinors in and enabling constructions like the via \mathfrak{su}(2) \cong \mathrm{so}(3). This indicator thus quantifies the "reality" of representations, with antilinear structures underpinning applications in and topological phases.

Properties

Algebraic Properties

The composition of two antilinear maps is a . Let S: V \to W and T: W \to U be antilinear maps between complex vector spaces. Then for any v \in V and \alpha \in \mathbb{C}, (T \circ S)(\alpha v) = T(\overline{\alpha} S(v)) = \alpha T(S(v)) = \alpha (T \circ S)(v), since the conjugate appears twice and cancels, yielding complex homogeneity. Additivity follows similarly from the additivity of each map. The kernel of an antilinear map T: V \to W is the set \ker T = \{ v \in V \mid T v = 0 \}, which forms a complex subspace of V. If T v = 0, then for \alpha \in \mathbb{C}, T(\alpha v) = \overline{\alpha} T v = 0, confirming closure under complex . The image \operatorname{im} T = \{ T v \mid v \in V \} is likewise a complex subspace of W. These subspaces satisfy the rank-nullity theorem over \mathbb{C}: \dim_{\mathbb{C}} \ker T + \dim_{\mathbb{C}} \operatorname{im} T = \dim_{\mathbb{C}} V, analogous to the linear case due to the underlying real-linearity structure. An antilinear T: V \to W between complex vector spaces admits an antilinear T^{-1}: W \to V. Bijectivity ensures a unique v = T^{-1} w for each w \in W with T v = w. For \beta \in \mathbb{C}, T(T^{-1}(\beta w)) = \beta w = \overline{\beta} T(T^{-1} w), implying T^{-1}(\beta w) = \overline{\beta} T^{-1} w, verifying antilinearity of the . Additivity holds by . The set \operatorname{Hom}_{\text{anti}}(V, W) of all antilinear maps from V to W forms a vector space over \mathbb{R}, with pointwise addition (S + T)(v) = S v + T v and real scalar multiplication (r S)(v) = r S v for r \in \mathbb{R}. This preserves antilinearity since \overline{r} = r. It is not closed under complex scalar multiplication, precluding a complex vector space structure. For finite-dimensional V and W with \dim_{\mathbb{C}} V = n and \dim_{\mathbb{C}} W = m, \dim_{\mathbb{R}} \operatorname{Hom}_{\text{anti}}(V, W) = 2 n m, matching the real dimension of the space of linear maps. Moreover, \operatorname{Hom}_{\text{anti}}(V, V) is a bimodule over the algebra of linear endomorphisms.

Analytic Properties

In normed vector spaces over the complex numbers, an antilinear map T: X \to Y is continuous if and only if it is bounded. Boundedness means there exists a constant M \geq 0 such that \|T(v)\| \leq M \|v\| for all v \in X, and the operator norm is given by \|T\| = \sup_{\|v\| \leq 1} \|T(v)\|. This equivalence follows from the same arguments as for linear maps, adapted to conjugate homogeneity, since |\bar{\alpha}| = |\alpha| preserves the necessary estimates in the proof. Furthermore, every bounded antilinear map between normed spaces is uniformly continuous, as the uniform bound on the difference \|T(v) - T(w)\| \leq \|T\| \|v - w\| holds directly. In Hilbert spaces, bounded antilinear operators are closely tied to sesquilinear forms. Specifically, for a bounded antilinear T: H \to H, the B(u, v) = \langle T u, v \rangle defines a that is antilinear in the first argument and linear in the second. Such forms can be recovered from their associated quadratic forms Q(w) = B(w, w) = \langle T w, w \rangle via the adapted for the complex case: B(u, v) = \frac{1}{4} \left[ Q(u + v) - Q(u - v) + i Q(u + i v) - i Q(u - i v) \right]. This identity holds for any sesquilinear form and ensures that the operator T is uniquely determined by the quadratic form on the diagonal. The spectrum of antilinear operators differs from that of linear ones due to the conjugate homogeneity. For eigenvalues, if T v = \lambda v for v \neq 0, then scaling v by a phase factor \mu = e^{i\theta} yields T (\mu v) = \bar{\mu} \lambda v = \mu^{-1} \lambda (\mu v), implying that the eigenvalues lie on circles centered at the origin in the complex plane rather than isolated points. In finite dimensions, an example is the spin-flip operator \theta_F on \mathbb{C}^2 defined by \theta_F (x, y) = (\bar{y}, -\bar{x}), which satisfies \theta_F^2 = -I and has an empty spectrum, as no eigenvectors exist.

Anti-Dual Space

Construction

The anti-dual space of a complex vector space V, denoted V^{\vee}, is defined as the set of all antilinear functionals \phi: V \to \mathbb{C}. This set consists of maps that are additive, \phi(u + v) = \phi(u) + \phi(v) for all u, v \in V, and conjugate homogeneous, \phi(\alpha v) = \bar{\alpha} \phi(v) for all \alpha \in \mathbb{C} and v \in V. The anti-dual space V^{\vee} forms a complex vector space under pointwise addition of functionals, (\phi + \psi)(v) = \phi(v) + \psi(v) for \phi, \psi \in V^{\vee} and v \in V, and scalar multiplication defined by (\alpha \cdot \phi)(v) = \alpha \, \phi(v) for \alpha \in \mathbb{C}, \phi \in V^{\vee}, and v \in V. This structure ensures that V^{\vee} is linear over \mathbb{C} in the usual sense, despite the antilinearity of each individual functional. A universal construction of the anti-dual space arises via the conjugate vector space \bar{V}, which is the set V equipped with the modified scalar multiplication \lambda \cdot v = \bar{\lambda} v for \lambda \in \mathbb{C} and v \in V. The (linear) dual space \bar{V}^* of \bar{V}, consisting of all \mathbb{C}-linear maps \bar{V} \to \mathbb{C}, is isomorphic to V^{\vee} as complex vector spaces. The explicit isomorphism sends a linear functional f \in \bar{V}^* to the antilinear functional \phi \in V^{\vee} given by \phi(v) = f(v) for all v \in V, since the linearity condition f(\bar{\lambda} v) = \lambda f(v) on \bar{V} corresponds precisely to the antilinearity \phi(\lambda v) = \bar{\lambda} \phi(v) on V. Alternatively, viewing the algebraic dual as a tensor product construction over the reals, V^{\vee} \cong V^* \otimes_{\mathbb{R}} \mathbb{C} as real modules, where V^* is the linear dual treated as a real vector space; the isomorphism extends the real-linear maps by complexification and selects the antilinear component. If V is finite-dimensional with \dim_{\mathbb{C}} V = n < \infty, then \dim_{\mathbb{C}} V^{\vee} = n. To see this, let \{v_1, \dots, v_n\} be a basis for V. Define the functionals \delta_i \in V^{\vee} by \delta_i\left( \sum_{j=1}^n \alpha_j v_j \right) = \bar{\alpha}_i for i = 1, \dots, n and \alpha_j \in \mathbb{C}. Each \delta_i is antilinear, and these form a basis for V^{\vee} because any antilinear functional is uniquely determined by its values on the basis (which can be arbitrary in \mathbb{C}), and follows from the fact that if \sum \beta_i \delta_i = 0, then evaluating on v_k yields \beta_k = 0.

Isomorphisms and Dualities

The anti-dual space V^\vee, consisting of antilinear functionals on a complex vector space V, is naturally isomorphic to the conjugate dual space \overline{V}^*, defined as the set of all functionals \psi: V \to \mathbb{C} satisfying \psi(\alpha v) = \overline{\alpha} \psi(v) for \alpha \in \mathbb{C} and v \in V. This identification arises because \overline{V}^* precisely captures the antilinear condition. The explicit isomorphism maps a linear functional \phi \in V^* (the standard dual space) to the antilinear functional \overline{\phi} \in V^\vee given by \overline{\phi}(v) = \overline{\phi(v)}. The inverse map sends \psi \in V^\vee to \overline{\psi} \in V^*, defined by \overline{\psi}(v) = \overline{\psi(v)}. Although this correspondence is conjugate-linear (i.e., \overline{\alpha \phi}(v) = \overline{\alpha} \cdot \overline{\phi}(v)), it establishes a bijective structure-preserving map between V^* and V^\vee as complex vector spaces, highlighting their structural equivalence. A key relation connects the anti-dual to the of the underlying real vector space V_\mathbb{R}. Specifically, the (V_\mathbb{R})^* \otimes_\mathbb{R} \mathbb{C} is isomorphic to V^* \oplus V^\vee as vector spaces, where (V_\mathbb{R})^* denotes the space of real-linear functionals on V_\mathbb{R}. This isomorphism extends real-linear functionals to the setting via conjugation in the scalar action, embedding antilinearity naturally as a direct summand. In finite dimensions, the isomorphism is and dimension-preserving (both sides have dimension $2 \dim_\mathbb{C} V). In dimensions, it holds algebraically but requires additional topological assumptions (e.g., ) for the continuous versions to align, underscoring the role of antilinear maps in bridging real and dualities without altering the underlying additive structure. For reflexive spaces, such as Hilbert spaces, biduality incorporates antilinear maps through natural identifications. The bidual (V^*)^* embeds V via the linear map, but in the Hilbert setting, the yields an antilinear isomorphism V \to V' (the continuous dual), extending to the bidual as V \cong (V')' with conjugate-linearity preserved in the pairing. The Hahn-Banach theorem, which extends linear functionals while preserving norms, adapts to antilinear functionals by applying it to the conjugate space \overline{V}, ensuring that bounded antilinear forms on subspaces extend to the whole space. This adaptation maintains reflexivity for the anti-dual structure, as the double anti-dual (V^\vee)^\vee identifies with V via linear maps (since composing two antilinear maps yields linearity), mirroring standard biduality but with conjugation in intermediate steps.

References

  1. [1]
    Numerical Ranges of Antilinear Operators | Integral Equations and ...
    May 21, 2024 · An antilinear operator acting on X is a bounded mapping satisfying and for all and . A typical example of an antilinear operator is the natural ...
  2. [2]
    [1507.06545] Anti- (Conjugate) Linearity - arXiv
    Jul 23, 2015 · After remembering elementary Tomita-Takesaki theory, antilinear maps, assiciated associated to a two-partite quantum system, are defined. By ...
  3. [3]
    [PDF] Appendix A Banach and Hilbert spaces - Texas A&M University
    7 (Linear, antilinear map). Let V, W be complex vector spaces. A map A : V → W is said to be linear if A(v1 + v2) ...
  4. [4]
    Analytic Continuation of Group Representations-V* - Project Euclid
    under the subgroup K by an n x n matrix representation σ. Then the cor- respondence defines an antilinear map H -> Γ(στ) that is an intertwining map with.
  5. [5]
    antilinear map in nLab
    ### Formal Definition of an Antilinear Map Between Complex Vector Spaces
  6. [6]
    Sesquilinear - an overview | ScienceDirect Topics
    Sesquilinear refers to a type of map that is linear in one argument and conjugate linear in the other, as illustrated by the complex sesquilinear map ...
  7. [7]
    [PDF] Anti- (Conjugate) Linearity - arXiv
    May 11, 2016 · This is an introduction to antilinear operators. In following E. P. Wigner the terminus “antilinear” is used as it is standard in Physics. ...
  8. [8]
    [PDF] arXiv:1211.6652v2 [math.QA] 5 Dec 2012
    Dec 5, 2012 · In Section 3 we defined antilinear maps and introduced complex conjugation of vector spaces as a way to turn antilinear maps into linear ones.
  9. [9]
    Real representations of C 2 -graded groups: The antilinear theory
    Feb 1, 2021 · We use the structure of finite-dimensional graded algebras to develop the theory of antilinear representations of finite C 2 -graded groups.
  10. [10]
  11. [11]
    [PDF] Implications of Time-Reversal Symmetry in Quantum Mechanics
    In quantum mechanics, the time reversal operator Θ acting on a state produces a state ... important properties of antilinear operators. Given an antilinear ...
  12. [12]
    [PDF] Antilinear superoperator, quantum geometric invariance, and ... - arXiv
    A crucial extreme example is the time reversal operator T, for which T2 = ±I. [6–9]. When T2 = −I, T must not be diagonalizable. An antilinear superoperator ...<|separator|>
  13. [13]
    [PDF] The Standard Model Parity, Charge Conjugation and Time Reversal
    The charge conjugation operation on the Dirac field must again interchange particles and anti-particles. The transformation therefore involves the hermitian.
  14. [14]
    [PDF] 7 Quantization of the Free Dirac Field
    7.1 The Dirac equation and Quantum Field Theory. The Dirac equation is a relativistic wave equation that describes the quan- tum dynamics of spinors.
  15. [15]
    [PDF] Introduction to representation theory - MIT Mathematics
    Jan 10, 2011 · 4.1 Frobenius-Schur indicator. Suppose that G is a finite group and V is an irreducible representation of G over C. We say that. V is. - of ...
  16. [16]
    [PDF] Representation theory
    The Frobenius-Schur indicator of an irreducible representation is +1 if it ... SO(3)∼=SU(2)/Z(SU(2)), the Schur multiplier of SO(3) is Z2, leading to two ...
  17. [17]
    [PDF] WRAP-Real-representations-C2-graded-antilinear-theory-Rumynin ...
    Jun 15, 2020 · Notice that the set of all A-homomorphisms HomA(V,W) is a real vector space. 2.2. Antilinear Maps and Matrices. We use the notation. ǫA = {. A ...
  18. [18]
    Chapter III Theory of Linear Operators in Hilbert Spaces
    A linear or antilinear transformation T of the normed space Jfx into the normed space ^V2 is continuous if and only if it is bounded. Proof. If || T || < + ...
  19. [19]
    [PDF] Chapter 8 Basics of Hermitian Geometry - UPenn CIS
    A Hermitian form is a sesquilinear form where '(v, u) = '(u, v). A Hermitian space is a pre-Hilbert space with a positive definite form.
  20. [20]
  21. [21]
    [PDF] Complex vector spaces, duals, and duels: Fun with a number, or two ...
    the complex linear operator on a complex vector space — is ...<|separator|>
  22. [22]
    [PDF] Dual space
    Mar 16, 2013 · Dual vector spaces defined on finite-dimensional vector spaces can be used for defining tensors.