Complemented subspace

Summary

In the branch of mathematics called functional analysis, a complemented subspace of a topological vector space is a vector subspace for which there exists some other vector subspace of called its (topological) complement in , such that is the direct sum in the category of topological vector spaces. Formally, topological direct sums strengthen the algebraic direct sum by requiring certain maps be continuous; the result retains many nice properties from the operation of direct sum in finite-dimensional vector spaces.

Every finite-dimensional subspace of a Banach space is complemented, but other subspaces may not. In general, classifying all complemented subspaces is a difficult problem, which has been solved only for some well-known Banach spaces.

The concept of a complemented subspace is analogous to, but distinct from, that of a set complement. The set-theoretic complement of a vector subspace is never a complementary subspace.

Preliminaries: definitions and notation edit

If   is a vector space and   and   are vector subspaces of   then there is a well-defined addition map

 
The map   is a morphism in the category of vector spaces — that is to say, linear.

Algebraic direct sum edit

The vector space   is said to be the algebraic direct sum (or direct sum in the category of vector spaces)   when any of the following equivalent conditions are satisfied:

  1. The addition map   is a vector space isomorphism.[1][2]
  2. The addition map is bijective.
  3.   and  ; in this case   is called an algebraic complement or supplement to   in   and the two subspaces are said to be complementary or supplementary.[2][3]

When these conditions hold, the inverse   is well-defined and can be written in terms of coordinates as

 
The first coordinate   is called the canonical projection of   onto  ; likewise the second coordinate is the canonical projection onto  [4]

Equivalently,   and   are the unique vectors in   and   respectively, that satisfy

 
As maps,
 
where   denotes the identity map on  .[2]

Motivation edit

Suppose that the vector space   is the algebraic direct sum of  . In the category of vector spaces, finite products and coproducts coincide: algebraically,   and   are indistinguishable. Given a problem involving elements of  , one can break the elements down into their components in   and  , because the projection maps defined above act as inverses to the natural inclusion of   and   into  . Then one can solve the problem in the vector subspaces and recombine to form an element of  .

In the category of topological vector spaces, that algebraic decomposition becomes less useful. The definition of a topological vector space requires the addition map   to be continuous; its inverse   may not be.[1] The categorical definition of direct sum, however, requires   and   to be morphisms — that is, continuous linear maps.

The space   is the topological direct sum of   and   if (and only if) any of the following equivalent conditions hold:

  1. The addition map   is a TVS-isomorphism (that is, a surjective linear homeomorphism).[1]
  2.   is the algebraic direct sum of   and   and also any of the following equivalent conditions:
    1. The inverse of the addition map   is continuous.
    2. Both canonical projections   and   are continuous.
    3. At least one of the canonical projections   and   is continuous.
    4. The canonical quotient map   is an isomorphism of topological vector spaces (i.e. a linear homeomorphism).[2]
  3.   is the direct sum of   and   in the category of topological vector spaces.
  4. The map   is bijective and open.
  5. When considered as additive topological groups,   is the topological direct sum of the subgroups   and  

The topological direct sum is also written  ; whether the sum is in the topological or algebraic sense is usually clarified through context.

Definition edit

Every topological direct sum is an algebraic direct sum  ; the converse is not guaranteed. Even if both   and   are closed in  ,   may still fail to be continuous.   is a (topological) complement or supplement to   if it avoids that pathology — that is, if, topologically,  . (Then   is likewise complementary to  .)[1] Condition 2(d) above implies that any topological complement of   is isomorphic, as a topological vector space, to the quotient vector space  .

  is called complemented if it has a topological complement   (and uncomplemented if not). The choice of   can matter quite strongly: every complemented vector subspace   has algebraic complements that do not complement   topologically.

Because a linear map between two normed (or Banach) spaces is bounded if and only if it is continuous, the definition in the categories of normed (resp. Banach) spaces is the same as in topological vector spaces.

Equivalent characterizations edit

The vector subspace   is complemented in   if and only if any of the following holds:[1]

  • There exists a continuous linear map   with image   such that  . That is,   is a continuous linear projection onto  . (In that case, algebraically  , and it is the continuity of   that implies that this is a complement.)
  • For every TVS   the restriction map   is surjective.[5]

If in addition   is Banach, then an equivalent condition is

  •   is closed in  , there exists another closed subspace  , and   is an isomorphism from the abstract direct sum   to  .

Examples edit

  • If   is a measure space and   has positive measure, then   is complemented in  .
  •  , the space of sequences converging to  , is complemented in  , the space of convergent sequences.
  • By Lebesgue decomposition,   is complemented in  .

Sufficient conditions edit

For any two topological vector spaces   and  , the subspaces   and   are topological complements in  .

Every algebraic complement of  , the closure of  , is also a topological complement. This is because   has the indiscrete topology, and so the algebraic projection is continuous.[6]

If   and   is surjective, then  .[2]

Finite dimension edit

Suppose   is Hausdorff and locally convex and   a free topological vector subspace: for some set  , we have   (as a t.v.s.). Then   is a closed and complemented vector subspace of  .[proof 1] In particular, any finite-dimensional subspace of   is complemented.[7]

In arbitrary topological vector spaces, a finite-dimensional vector subspace   is topologically complemented if and only if for every non-zero  , there exists a continuous linear functional on   that separates   from  .[1] For an example in which this fails, see § Fréchet spaces.

Finite codimension edit

Not all finite-codimensional vector subspaces of a TVS are closed, but those that are, do have complements.[7][8]

Hilbert spaces edit

In a Hilbert space, the orthogonal complement   of any closed vector subspace   is always a topological complement of  . This property characterizes Hilbert spaces within the class of Banach spaces: every infinite dimensional, non-Hilbert Banach space contains a closed uncomplemented subspace, a deep theorem of Joram Lindenstrauss and Lior Tzafriri.[9][3]

Fréchet spaces edit

Let   be a Fréchet space over the field  . Then the following are equivalent:[10]

  1.   is not normable (that is, any continuous norm does not generate the topology)
  2.   contains a vector subspace TVS-isomorphic to  
  3.   contains a complemented vector subspace TVS-isomorphic to  .

Properties; examples of uncomplemented subspaces edit

A complemented (vector) subspace of a Hausdorff space   is necessarily a closed subset of  , as is its complement.[1][proof 2]

From the existence of Hamel bases, every infinite-dimensional Banach space contains unclosed linear subspaces.[proof 3] Since any complemented subspace is closed, none of those subspaces is complemented.

Likewise, if   is a complete TVS and   is not complete, then   has no topological complement in  [11]

Applications edit

If   is a continuous linear surjection, then the following conditions are equivalent:

  1. The kernel of   has a topological complement.
  2. There exists a "right inverse": a continuous linear map   such that  , where   is the identity map.[5]

(Note: This claim is an erroneous exercise given by Trèves. Let   and   both be   where   is endowed with the usual topology, but   is endowed with the trivial topology. The identity map   is then a continuous, linear bijection but its inverse is not continuous, since   has a finer topology than  . The kernel   has   as a topological complement, but we have just shown that no continuous right inverse can exist. If   is also open (and thus a TVS homomorphism) then the claimed result holds.)

The Method of Decomposition edit

Topological vector spaces admit the following Cantor-Schröder-Bernstein–type theorem:

Let   and   be TVSs such that   and   Suppose that   contains a complemented copy of   and   contains a complemented copy of   Then   is TVS-isomorphic to  

The "self-splitting" assumptions that   and   cannot be removed: Tim Gowers showed in 1996 that there exist non-isomorphic Banach spaces   and  , each complemented in the other.[12]

In classical Banach spaces edit

Understanding the complemented subspaces of an arbitrary Banach space   up to isomorphism is a classical problem that has motivated much work in basis theory, particularly the development of absolutely summing operators. The problem remains open for a variety of important Banach spaces, most notably the space  .[13]

For some Banach spaces the question is closed. Most famously, if   then the only complemented infinite-dimensional subspaces of   are isomorphic to   and the same goes for   Such spaces are called prime (when their only infinite-dimensional complemented subspaces are isomorphic to the original). These are not the only prime spaces, however.[13]

The spaces   are not prime whenever   in fact, they admit uncountably many non-isomorphic complemented subspaces.[13]

The spaces   and   are isomorphic to   and   respectively, so they are indeed prime.[13]

The space   is not prime, because it contains a complemented copy of  . No other complemented subspaces of   are currently known.[13]

Indecomposable Banach spaces edit

An infinite-dimensional Banach space is called indecomposable whenever its only complemented subspaces are either finite-dimensional or -codimensional. Because a finite-codimensional subspace of a Banach space   is always isomorphic to   indecomposable Banach spaces are prime.

The most well-known example of indecomposable spaces are in fact hereditarily indecomposable, which means every infinite-dimensional subspace is also indecomposable.[14]

See also edit

Proofs edit

  1. ^   is closed because   is complete and   is Hausdorff.

    Let   be a TVS-isomorphism; each   is a continuous linear functional. By the Hahn–Banach theorem, we may extend each   to a continuous linear functional   on   The joint map   is a continuous linear surjection whose restriction to   is  . The composition   is then a continuous continuous projection onto  .

  2. ^ In a Hausdorff space,   is closed. A complemented space is the kernel of the (continuous) projection onto its complement. Thus it is the preimage of   under a continuous map, and so closed.
  3. ^ Any sequence   defines a summation map  . But if   are (algebraically) linearly independent and   has full support, then  .

References edit

  1. ^ a b c d e f g Grothendieck 1973, pp. 34–36.
  2. ^ a b c d e Fabian, Marián J.; Habala, Petr; Hájek, Petr; Montesinos Santalucía, Vicente; Zizler, Václav (2011). Banach Space Theory: The Basis for Linear and Nonlinear Analysis (PDF). New York: Springer. pp. 179–181. doi:10.1007/978-1-4419-7515-7. ISBN 978-1-4419-7515-7.
  3. ^ a b Brezis, Haim (2011). Functional Analysis, Sobolev Spaces, and Partial Differential Equations. Universitext. New York: Springer. pp. 38–39. ISBN 978-0-387-70913-0.
  4. ^ Schaefer & Wolff 1999, pp. 19–24.
  5. ^ a b Trèves 2006, p. 36.
  6. ^ Wilansky 2013, p. 63.
  7. ^ a b Rudin 1991, p. 106.
  8. ^ Serre, Jean-Pierre (1955). "Un théoreme de dualité". Commentarii Mathematici Helvetici. 29 (1): 9–26. doi:10.1007/BF02564268. S2CID 123643759.
  9. ^ Lindenstrauss, J., & Tzafriri, L. (1971). On the complemented subspaces problem. Israel Journal of Mathematics, 9, 263-269.
  10. ^ Jarchow 1981, pp. 129–130.
  11. ^ Schaefer & Wolff 1999, pp. 190–202.
  12. ^ Narici & Beckenstein 2011, pp. 100–101.
  13. ^ a b c d e Albiac, Fernando; Kalton, Nigel J. (2006). Topics in Banach Space Theory. GTM 233 (2nd ed.). Switzerland: Springer (published 2016). pp. 29–232. doi:10.1007/978-3-319-31557-7. ISBN 978-3-319-31557-7.
  14. ^ Argyros, Spiros; Tolias, Andreas (2004). Methods in the Theory of Hereditarily Indecomposable Banach Spaces. American Mathematical Soc. ISBN 978-0-8218-3521-0.

Bibliography edit

  • Bachman, George; Narici, Lawrence (2000). Functional Analysis (Second ed.). Mineola, New York: Dover Publications. ISBN 978-0486402512. OCLC 829157984.
  • Grothendieck, Alexander (1973). Topological Vector Spaces. Translated by Chaljub, Orlando. New York: Gordon and Breach Science Publishers. ISBN 978-0-677-30020-7. OCLC 886098.
  • Jarchow, Hans (1981). Locally convex spaces. Stuttgart: B.G. Teubner. ISBN 978-3-519-02224-4. OCLC 8210342.
  • Narici, Lawrence; Beckenstein, Edward (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Rudin, Walter (1991). Functional Analysis. International Series in Pure and Applied Mathematics. Vol. 8 (Second ed.). New York, NY: McGraw-Hill Science/Engineering/Math. ISBN 978-0-07-054236-5. OCLC 21163277.
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. Vol. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.
  • Trèves, François (2006) [1967]. Topological Vector Spaces, Distributions and Kernels. Mineola, N.Y.: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.
  • Wilansky, Albert (2013). Modern Methods in Topological Vector Spaces. Mineola, New York: Dover Publications, Inc. ISBN 978-0-486-49353-4. OCLC 849801114.