Sesquilinear form




In mathematics, a sesquilinear form is a generalization of a bilinear form that, in turn, is a generalization of the concept of the dot product of Euclidean space. A bilinear form is linear in each of its arguments, but a sesquilinear form allows one of the arguments to be "twisted" in a semilinear manner, thus the name; which originates from the Latin numerical prefix sesqui- meaning "one and a half". The basic concept of the dot product – producing a scalar from a pair of vectors – can be generalized by allowing a broader range of scalar values and, perhaps simultaneously, by widening the definition of what a vector is.


A motivating special case is a sesquilinear form on a complex vector space, V. This is a map V × VC that is linear in one argument and "twists" the linearity of the other argument by complex conjugation (referred to as being antilinear in the other argument). This case arises naturally in mathematical physics applications. Another important case allows the scalars to come from any field and the twist is provided by a field automorphism.


An application in projective geometry requires that the scalars come from a division ring (skewfield), K, and this means that the "vectors" should be replaced by elements of a K-module. In a very general setting, sesquilinear forms can be defined over R-modules for arbitrary rings R.




Contents






  • 1 Convention


  • 2 Complex vector spaces


    • 2.1 Matrix representation


    • 2.2 Hermitian form


    • 2.3 Skew-Hermitian form




  • 3 Over a division ring


    • 3.1 Definition


    • 3.2 Orthogonality


    • 3.3 Reflexivity


    • 3.4 Hermitian variations


    • 3.5 Example




  • 4 In projective geometry


  • 5 Over arbitrary rings


  • 6 See also


  • 7 Notes


  • 8 References


  • 9 External links





Convention


Conventions differ as to which argument should be linear. In the commutative case, we shall take the first to be linear, as is common in the mathematical literature, except in the section devoted to sesquilinear forms on complex vector spaces. There we use the other convention and take the first argument to be conjugate-linear (i.e. antilinear) and the second to be linear. This is the convention used mostly by mathematical physicists[1] and originates in Dirac's bra–ket notation in quantum mechanics.


In the more general noncommutative setting, with right modules we take the second argument to be linear and with left modules we take the first argument to be linear.



Complex vector spaces


Over a complex vector space V a map φ : V × VC is sesquilinear if


φ(x+y,z+w)=φ(x,z)+φ(x,w)+φ(y,z)+φ(y,w)φ(ax,by)=a¯(x,y){displaystyle {begin{aligned}&varphi (x+y,z+w)=varphi (x,z)+varphi (x,w)+varphi (y,z)+varphi (y,w)\&varphi (ax,by)={overline {a}}b,varphi (x,y)end{aligned}}}{begin{aligned}&varphi (x+y,z+w)=varphi (x,z)+varphi (x,w)+varphi (y,z)+varphi (y,w)\&varphi (ax,by)={overline {a}}b,varphi (x,y)end{aligned}}

for all x, y, z, wV and all a, bC. a is the complex conjugate of a.


A complex sesquilinear form can also be viewed as a complex bilinear map


×V→C{displaystyle {overline {V}}times Vto mathbf {C} }{overline {V}}times Vto mathbf {C}

where V is the complex conjugate vector space to V. By the universal property of tensor products these are in one-to-one correspondence with complex linear maps


V→C.{displaystyle {overline {V}}otimes Vto mathbf {C} .}{overline {V}}otimes Vto mathbf {C} .

For a fixed z in V the map wφ(z, w) is a linear functional on V (i.e. an element of the dual space V). Likewise, the wφ(w, z) is a conjugate-linear functional on V.


Given any complex sesquilinear form φ on V we can define a second complex sesquilinear form ψ via the conjugate transpose:


ψ(w,z)=φ(z,w)¯.{displaystyle psi (w,z)={overline {varphi (z,w)}}.}psi (w,z)={overline {varphi (z,w)}}.

In general, ψ and φ will be different. If they are the same then φ is said to be Hermitian. If they are negatives of one another, then φ is said to be skew-Hermitian. Every sesquilinear form can be written as a sum of a Hermitian form and a skew-Hermitian form.



Matrix representation


If V is a finite-dimensional complex vector space, then relative to any basis { ei } of V, a sesquilinear form is represented by a matrix Φ, w by the column vector w, and z by the column vector z:


φ(w,z)=φ(∑iwiei,∑jzjej)=∑i∑jwi¯zjφ(ei,ej)=w¯z.{displaystyle varphi (w,z)=varphi left(sum _{i}w_{i}e_{i},sum _{j}z_{j}e_{j}right)=sum _{i}sum _{j}{overline {w_{i}}}z_{j}varphi (e_{i},e_{j})={overline {mathbf {w} }}^{mathrm {T} }mathbf {Phi } mathbf {z} .}{displaystyle varphi (w,z)=varphi left(sum _{i}w_{i}e_{i},sum _{j}z_{j}e_{j}right)=sum _{i}sum _{j}{overline {w_{i}}}z_{j}varphi (e_{i},e_{j})={overline {mathbf {w} }}^{mathrm {T} }mathbf {Phi } mathbf {z} .}

The components of Φ are given by Φij = φ(ei, ej).



Hermitian form


The term Hermitian form may also refer to a different concept than that explained below: it may refer to a certain differential form on a Hermitian manifold.

A complex Hermitian form (also called a symmetric sesquilinear form), is a sesquilinear form h : V × VC such that


h(w,z)=h(z,w)¯.{displaystyle h(w,z)={overline {h(z,w)}}.}h(w,z)={overline {h(z,w)}}.

The standard Hermitian form on Cn is given (again, using the "physics" convention of linearity in the second and conjugate linearity in the first variable) by


w,z⟩=∑i=1nwi¯zi.{displaystyle langle w,zrangle =sum _{i=1}^{n}{overline {w_{i}}}z_{i}.}langle w,zrangle =sum _{i=1}^{n}{overline {w_{i}}}z_{i}.

More generally, the inner product on any complex Hilbert space is a Hermitian form.


A vector space with a Hermitian form (V, h) is called a Hermitian space.


The matrix representation of a complex Hermitian form is a Hermitian matrix.


A complex Hermitian form applied to a single vector


|z|h=h(z,z){displaystyle |z|_{h}=h(z,z)}{displaystyle |z|_{h}=h(z,z)}

is always real. One can show that a complex sesquilinear form is Hermitian iff the associated quadratic form is real for all zV.



Skew-Hermitian form


A complex skew-Hermitian form (also called an antisymmetric sesquilinear form), is a complex sesquilinear form s : V × VC such that


s(w,z)=−s(z,w)¯.{displaystyle s(w,z)=-{overline {s(z,w)}}.}s(w,z)=-{overline {s(z,w)}}.

Every complex skew-Hermitian form can be written as i times a Hermitian form.


The matrix representation of a complex skew-Hermitian form is a skew-Hermitian matrix.


A complex skew-Hermitian form applied to a single vector


|z|s=s(z,z){displaystyle |z|_{s}=s(z,z)}{displaystyle |z|_{s}=s(z,z)}

is always pure imaginary.



Over a division ring


This section applies unchanged when the division ring K is commutative. More specific terminology then also applies: the division ring is a field, the anti-automorphism is also an automorphism, and the right module is a vector space. The following applies to a left module with suitable reordering of expressions.



Definition


A σ-sesquilinear form over a right K-module M is a bi-additive map φ : M × MK with an associated anti-automorphism σ of a division ring K such that, for all x, yM and all α, βK,


φ(xα,yβ)=σ(x,y)β.{displaystyle varphi (xalpha ,ybeta )=sigma (alpha ),varphi (x,y),beta .}{displaystyle varphi (xalpha ,ybeta )=sigma (alpha ),varphi (x,y),beta .}

The associated anti-automorphism σ for any nonzero sesquilinear form φ is uniquely determined by φ.



Orthogonality


Given a sesquilinear form φ over a module M and a subspace W of M, the orthogonal complement of W with respect to φ is


W⊥={v∈M∣φ(v,w)=0, ∀w∈W}.{displaystyle W^{perp }={mathbf {v} in Mmid varphi (mathbf {v} ,mathbf {w} )=0, forall mathbf {w} in W}.}{displaystyle W^{perp }={mathbf {v} in Mmid varphi (mathbf {v} ,mathbf {w} )=0, forall mathbf {w} in W}.}

Similarly, xM is orthogonal to yM with respect to φ, written xφy (or simply xy if φ can be inferred from the context), when φ(x, y) = 0. This relation need not be symmetric, i.e. xy does not imply yx (but see § Reflexivity below).



Reflexivity


A sesquilinear form φ is reflexive if, for all x, yM,



φ(x,y)=0{displaystyle varphi (x,y)=0}{displaystyle varphi (x,y)=0} implies φ(y,x)=0.{displaystyle varphi (y,x)=0.}{displaystyle varphi (y,x)=0.}

That is, a sesquilinear form is reflexive precisely when the derived orthogonality relation is symmetric.



Hermitian variations


A σ-sesquilinear form φ is called (σ, ε)-Hermitian if there exists εK such that, for all x, yM,


φ(x,y)=σ(y,x))ε.{displaystyle varphi (x,y)=sigma (varphi (y,x)),varepsilon .}{displaystyle varphi (x,y)=sigma (varphi (y,x)),varepsilon .}

If ε = 1, the form is called σ-Hermitian, and if ε = −1, it is called σ-anti-Hermitian. (When σ is implied, respectively simply Hermitian or anti-Hermitian.)


For a nonzero (σ, ε)-Hermitian form, it follows that, for all αK,



σ)=ε1{displaystyle sigma (varepsilon )=varepsilon ^{-1}}{displaystyle sigma (varepsilon )=varepsilon ^{-1}}

σ))=εαε1.{displaystyle sigma (sigma (alpha ))=varepsilon alpha varepsilon ^{-1}.}{displaystyle sigma (sigma (alpha ))=varepsilon alpha varepsilon ^{-1}.}


It also follows that φ(x, x) is a fixed point of the map ασ(α)ε. The fixed points of this map from a subgroup of the additive group of K.


A (σ, ε)-Hermitian form is reflexive, and every reflexive σ-sesquilinear form is (σ, ε)-Hermitian for some ε.[2][3][4][5]


In the special case that σ is the identity map (i.e., σ = id), K is commutative, φ is a bilinear form and ε2 = 1. Then for ε = 1 the bilinear form is called symmetric, and for ε = −1 is called skew-symmetric.[6]



Example


Let V be the three dimensional vector space over the finite field F = GF(q2), where q is a prime power. With respect to the standard basis we can write x = (x1, x2, x3) and y = (y1, y2, y3) and define the map φ by:


φ(x,y)=x1y1q+x2y2q+x3y3q.{displaystyle varphi (x,y)=x_{1}y_{1}{}^{q}+x_{2}y_{2}{}^{q}+x_{3}y_{3}{}^{q}.}varphi (x,y)=x_{1}y_{1}{}^{q}+x_{2}y_{2}{}^{q}+x_{3}y_{3}{}^{q}.

The map σ : ttq is an involutory automorphism of F. The map φ is then a σ-sesquilinear form. The matrix Mφ associated to this form is the identity matrix. This is a Hermitian form.



In projective geometry


In a projective geometry G, a permutation δ of the subspaces that inverts inclusion, i.e.



STTδSδ for all subspaces S, T of G,

is called a correlation. A result of Birkhoff and von Neumann (1936)[7] shows that the correlations of desarguesian projective geometries correspond to the nondegenerate sesquilinear forms on the underlying vector space.[5] A sesquilinear form φ is nondegenerate if φ(x, y) = 0 for all y in V (if and) only if x = 0.


To achieve full generality of this statement, and since every desarguesian projective geometry may be coordinatized by a division ring, Reinhold Baer extended the definition of a sesquilinear form to a division ring, which requires replacing vector spaces by R-modules.[8] (In the geometric literature these are still referred to as either left or right vector spaces over skewfields.)[9]



Over arbitrary rings


The specialization of the above section to skewfields was a consequence of the application to projective geometry, and not intrinsic to the nature of sesquilinear forms. Only the minor modifications needed to take into account the non-commutativity of multiplication are required to generalize the arbitrary field version of the definition to arbitrary rings.


Let R be a ring, V an R-module and σ an antiautomorphism of R.


A map φ : V × VR is σ-sesquilinear if



φ(x+y,z+w)=φ(x,z)+φ(x,w)+φ(y,z)+φ(y,w){displaystyle varphi (x+y,z+w)=varphi (x,z)+varphi (x,w)+varphi (y,z)+varphi (y,w)}{displaystyle varphi (x+y,z+w)=varphi (x,z)+varphi (x,w)+varphi (y,z)+varphi (y,w)}

φ(cx,dy)=cφ(x,y)σ(d){displaystyle varphi (cx,dy)=c,varphi (x,y),sigma (d)}{displaystyle varphi (cx,dy)=c,varphi (x,y),sigma (d)}


for all x, y, z, wV and all c, dR.


A element x is orthogonal to another element y with respect to the sesquilinear form φ (written xy) if φ(x, y) = 0. This relation need not be symmetric, i.e. xy does not imply yx.


A sesquilinear form φ : V × VR is reflexive (or orthosymmetric) if φ(x, y) = 0 implies φ(y, x) = 0 for all x, yV.


A sesquilinear form φ : V × VR is Hermitian if there exists σ such that[10]:325


φ(x,y)=σ(y,x)){displaystyle varphi (x,y)=sigma (varphi (y,x))}{displaystyle varphi (x,y)=sigma (varphi (y,x))}

for all x, yV. A Hermitian form is necessarily reflexive, and if it is nonzero, the associated antiautomorphism σ is an involution (i.e. of order 2).


Since for an antiautomorphism σ we have σ(st) = σ(t) σ(s) for all s, t in R, if σ = id, then R must be commutative and φ is a bilinear form. In particular, if, in this case, R is a skewfield, then R is a field and V is a vector space with a bilinear form.


An antiautomorphism σ : RR can also be viewed as an isomorphism of RRop, the opposite ring based on the same set with the same addition, but whose multiplication operation () is defined by ab = ba, where the product on the right is the product in R. It follows from this that a right (left) R-module V can be turned into a left (right) Rop-module, Vo.[11] Thus, the sesquilinear form φ : V × VR can be viewed as a bilinear form φ′ : V × VoR.



See also


  • *-ring


Notes





  1. ^ footnote 1 in Anthony Knapp Basic Algebra (2007) pg. 255


  2. ^
    "Combinatorics", Proceedings of the NATO Advanced Study Institute, Held at Nijenrode Castle, Breukelen, The Netherlands, 8–20 July 1974, D. Reidel: 456–457, 1975.mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em} – [1]



  3. ^
    Sesquilinear form at EOM



  4. ^ Simeon Ball (2015), Finite Geometry and Combinatorial Applications, Cambridge University Press, p. 28 – [2]


  5. ^ ab
    Dembowski 1968, p. 42



  6. ^ When char K = 2, skew-symmetric and symmetric bilinear forms coincide since then 1 = −1. In all cases, alternating bilinear forms are a subset of skew-symmetric bilinear forms, and need not be considered separately.


  7. ^ Birkhoff, G.; von Neumann, J. (1936), "The logic of quantum mechanics", Annals of Mathematics, 37: 823–843, doi:10.2307/1968621


  8. ^ Baer, Reinhold (2005) [1952], Linear Algebra and Projective Geometry, Dover, ISBN 978-0-486-44565-6


  9. ^ Baer's terminology gives a third way to refer to these ideas, so he must be read with caution.


  10. ^ Faure, Claude-Alain; Frölicher, Alfred (2000), Modern Projective Geometry, Kluwer Academic Publishers


  11. ^ Jacobson 2009, p. 164




References




  • Dembowski, Peter (1968), Finite geometries, Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 44, Berlin, New York: Springer-Verlag, ISBN 3-540-61786-8, MR 0233275


  • Gruenberg, K.W.; Weir, A.J. (1977), Linear Geometry (2nd ed.), Springer, ISBN 0-387-90227-9


  • Jacobson, Nathan J. (2009) [1985], Basic Algebra I (2nd ed.), Dover, ISBN 978-0-486-47189-1



External links



  • Hazewinkel, Michiel, ed. (2001) [1994], "Sesquilinear form", Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4



這個網誌中的熱門文章

Tangent Lines Diagram Along Smooth Curve

Yusuf al-Mu'taman ibn Hud

Zucchini