Movatterモバイル変換


[0]ホーム

URL:


Jump to content
WikipediaThe Free Encyclopedia
Search

Open mapping theorem (functional analysis)

From Wikipedia, the free encyclopedia
Condition for a linear operator to be open

This article is about open mapping theorem infunctional analysis. For other results with the same name, seeOpen mapping theorem.

Infunctional analysis, theopen mapping theorem, also known as theBanach–Schauder theorem or theBanach theorem[1] (named afterStefan Banach andJuliusz Schauder), is a fundamental result that states that if abounded orcontinuous linear operator betweenBanach spaces issurjective then it is anopen map.

A special case is also called thebounded inverse theorem (also called inverse mapping theorem or Banach isomorphism theorem), which states that a bijective bounded linear operatorT{\displaystyle T} from one Banach space to another has bounded inverseT1{\displaystyle T^{-1}}.

Statement and proof

[edit]

Open mapping theorem[2][3] LetT:EF{\displaystyle T:E\to F} be a surjective continuous linear map between Banach spaces (or more generallyFréchet spaces). ThenT{\displaystyle T} is an open mapping (that is, ifUE{\displaystyle U\subset E} is an open subset, thenT(U){\displaystyle T(U)} is open).

The proof here uses theBaire category theorem, andcompleteness of bothE{\displaystyle E} andF{\displaystyle F} is essential to the theorem. The statement of the theorem is no longer true if either space is assumed to be only anormed vector space; see§ Counterexample.

The proof is based on the following lemmas, which are also somewhat of independent interest. A linear mapf:EF{\displaystyle f:E\to F} between topological vector spaces is said to benearly open if, for each neighborhoodU{\displaystyle U} of zero, the closuref(U)¯{\displaystyle {\overline {f(U)}}} contains a neighborhood of zero. The next lemma may be thought of as a weak version of the open mapping theorem.

Lemma[4][5] A linear mapf:EF{\displaystyle f:E\to F} between normed spaces is nearly open if the image off{\displaystyle f} isnon-meager inF{\displaystyle F}. (The continuity is not needed.)

Proof: ShrinkingU{\displaystyle U}, we can assumeU{\displaystyle U} is an open ball centered at zero. We havef(E)=f(nNnU)=nNf(nU){\displaystyle f(E)=f\left(\bigcup _{n\in \mathbb {N} }nU\right)=\bigcup _{n\in \mathbb {N} }f(nU)}. Thus, somef(nU)¯{\displaystyle {\overline {f(nU)}}} contains an interior pointy{\displaystyle y}; that is, for some radiusr>0{\displaystyle r>0},

B(y,r)f(nU)¯.{\displaystyle B(y,r)\subset {\overline {f(nU)}}.}

Then for anyv{\displaystyle v} inF{\displaystyle F} withv<r{\displaystyle \|v\|<r}, by linearity, convexity and(1)UU{\displaystyle (-1)U\subset U},

v=vy+yf(nU)¯+f(nU)¯f(2nU)¯{\displaystyle v=v-y+y\in {\overline {f(-nU)}}+{\overline {f(nU)}}\subset {\overline {f(2nU)}}},

which proves the lemma by dividing by2n{\displaystyle 2n}.{\displaystyle \square } (The same proof works ifE,F{\displaystyle E,F} are pre-Fréchet spaces.)

The completeness on the domain then allows to upgrade nearly open to open.

Lemma (Schauder)[6][7] Letf:EF{\displaystyle f:E\to F} be a continuous linear map between normed spaces.

Iff{\displaystyle f} is nearly-open and ifE{\displaystyle E} is complete, thenf{\displaystyle f} is open and surjective.

More precisely, ifB(0,δ)f(B(0,1))¯{\displaystyle B(0,\delta )\subset {\overline {f(B(0,1))}}} for someδ>0{\displaystyle \delta >0} and ifE{\displaystyle E} is complete, then

B(0,δ)f(B(0,1)){\displaystyle B(0,\delta )\subset f(B(0,1))}

whereB(x,r){\displaystyle B(x,r)} is an open ball with radiusr{\displaystyle r} and centerx{\displaystyle x}.

Proof: Lety{\displaystyle y} be inB(0,δ){\displaystyle B(0,\delta )} andcn>0{\displaystyle c_{n}>0} some sequence. We have:B(0,δ)¯f(B(0,1))¯{\displaystyle {\overline {B(0,\delta )}}\subset {\overline {f(B(0,1))}}}. Thus, for eachϵ>0{\displaystyle \epsilon >0} andz{\displaystyle z} inF{\displaystyle F}, we can find anx{\displaystyle x} withx<δ1z{\displaystyle \|x\|<\delta ^{-1}\|z\|} andz{\displaystyle z} inB(f(x),ϵ){\displaystyle B(f(x),\epsilon )}. Thus, takingz=y{\displaystyle z=y}, we find anx1{\displaystyle x_{1}} such that

yf(x1)<c1,x1<δ1y.{\displaystyle \|y-f(x_{1})\|<c_{1},\,\|x_{1}\|<\delta ^{-1}\|y\|.}

Applying the same argument withz=yf(x1){\displaystyle z=y-f(x_{1})}, we then find anx2{\displaystyle x_{2}} such that

yf(x1)f(x2)<c2,x2<δ1c1{\displaystyle \|y-f(x_{1})-f(x_{2})\|<c_{2},\,\|x_{2}\|<\delta ^{-1}c_{1}}

where we observedx2<δ1z<δ1c1{\displaystyle \|x_{2}\|<\delta ^{-1}\|z\|<\delta ^{-1}c_{1}}. Then so on. Thus, ifc:=cn<{\displaystyle c:=\sum c_{n}<\infty }, we found a sequencexn{\displaystyle x_{n}} such thatx=1xn{\displaystyle x=\sum _{1}^{\infty }x_{n}} converges andf(x)=y{\displaystyle f(x)=y}. Also,

x1xnδ1y+δ1c.{\displaystyle \|x\|\leq \sum _{1}^{\infty }\|x_{n}\|\leq \delta ^{-1}\|y\|+\delta ^{-1}c.}

Sinceδ1y<1{\displaystyle \delta ^{-1}\|y\|<1}, by makingc{\displaystyle c} small enough, we can achievex<1{\displaystyle \|x\|<1}.{\displaystyle \square } (Again the same proof is valid ifE,F{\displaystyle E,F} are pre-Fréchet spaces.)

Proof of the theorem: By Baire's category theorem, the first lemma applies. Then the conclusion of the theorem follows from the second lemma.{\displaystyle \square }

In general, a continuous bijection between topological spaces is not necessarily a homeomorphism. The open mapping theorem, when it applies, implies the bijectivity is enough:

Corollary (Bounded inverse theorem)[8] A continuous bijective linear operator between Banach spaces (or Fréchet spaces) has continuous inverse. That is, the inverse operator is continuous.

Even though the above bounded inverse theorem is a special case of the open mapping theorem, the open mapping theorem in turn follows from that. Indeed, a surjective continuous linear operatorT:EF{\displaystyle T:E\to F} factors as

T:EpE/kerTT0F.{\displaystyle T:E{\overset {p}{\to }}E/\operatorname {ker} T{\overset {T_{0}}{\to }}F.}

Here,T0{\displaystyle T_{0}} is continuous and bijective and thus is a homeomorphism by the bounded inverse theorem; in particular, it is an open mapping. As a quotient map for topological groups is open,T{\displaystyle T} is open then.

Because the open mapping theorem and the bounded inverse theorem are essentially the same result, they are often simply calledBanach's theorem.

Transpose formulation

[edit]

Here is a formulation of the open mapping theorem in terms of thetranspose of an operator.

Theorem[6]LetX{\displaystyle X} andY{\displaystyle Y} be Banach spaces, letBX{\displaystyle B_{X}} andBY{\displaystyle B_{Y}} denote their open unit balls, and letT:XY{\displaystyle T:X\to Y} be a bounded linear operator.Ifδ>0{\displaystyle \delta >0} then among the following four statements we have(1)(2)(3)(4){\displaystyle (1)\implies (2)\implies (3)\implies (4)} (with the sameδ{\displaystyle \delta })

  1. δyTy{\displaystyle \delta \left\|y'\right\|\leq \left\|T'y'\right\|} for allyY{\displaystyle y'\in Y'} = continuous dual ofY{\displaystyle Y};
  2. δBYT(BX)¯{\displaystyle \delta B_{Y}\subset {\overline {T\left(B_{X}\right)}}};
  3. δBYT(BX){\displaystyle \delta B_{Y}\subset {T\left(B_{X}\right)}};
  4. T{\displaystyle T} is surjective.

Furthermore, ifT{\displaystyle T} is surjective then (1) holds for someδ>0.{\displaystyle \delta >0.}

Proof: The idea of 1.{\displaystyle \Rightarrow } 2. is to show:yT(BX)¯y>δ,{\displaystyle y\notin {\overline {T(B_{X})}}\Rightarrow \|y\|>\delta ,} and that follows from theHahn–Banach theorem. 2.{\displaystyle \Rightarrow } 3. is exactly the second lemma in§ Statement and proof. Finally, 3.{\displaystyle \Rightarrow } 4. is trivial and 4.{\displaystyle \Rightarrow } 1. easily follows from the open mapping theorem.{\displaystyle \square }

Alternatively, 1. implies thatT{\displaystyle T'} is injective and has closed image and then by theclosed range theorem, that impliesT{\displaystyle T} has dense image and closed image, respectively; i.e.,T{\displaystyle T} is surjective. Hence, the above result is a variant of a special case of the closed range theorem.

Quantative formulation

[edit]

Terence Tao gives the following quantitative formulation of the theorem:[9]

TheoremLetT:EF{\displaystyle T:E\to F} be a bounded operator between Banach spaces. Then the following are equivalent:

  1. T{\displaystyle T} is open.
  2. T{\displaystyle T} is surjective.
  3. There exists a constantC>0{\displaystyle C>0} such that, for eachf{\displaystyle f} inF{\displaystyle F}, the equationTu=f{\displaystyle Tu=f} has a solutionu{\displaystyle u} withuCf{\displaystyle \|u\|\leq C\|f\|}.
  4. 3. holds forf{\displaystyle f} in some dense subspace ofF{\displaystyle F}.

The proof follows acycle of implications14321{\displaystyle 1\Rightarrow 4\Rightarrow 3\Rightarrow 2\Rightarrow 1}. Here21{\displaystyle 2\Rightarrow 1} is the usual open mapping theorem.

14{\displaystyle 1\Rightarrow 4}: For somer>0{\displaystyle r>0}, we haveB(0,2)T(B(0,r)){\displaystyle B(0,2)\subset T(B(0,r))} whereB{\displaystyle B} means an open ball. Thenff=T(uf){\displaystyle {\frac {f}{\|f\|}}=T\left({\frac {u}{\|f\|}}\right)} for someuf{\displaystyle {\frac {u}{\|f\|}}} inB(0,r){\displaystyle B(0,r)}. That is,Tu=f{\displaystyle Tu=f} withu<rf{\displaystyle \|u\|<r\|f\|}.

43{\displaystyle 4\Rightarrow 3}: We can writef=0fj{\displaystyle f=\sum _{0}^{\infty }f_{j}} withfj{\displaystyle f_{j}} in the dense subspace and the sum converging in norm. Then, sinceE{\displaystyle E} is complete,u=0uj{\displaystyle u=\sum _{0}^{\infty }u_{j}} withujCfj{\displaystyle \|u_{j}\|\leq C\|f_{j}\|} andTuj=fj{\displaystyle Tu_{j}=f_{j}} is a required solution.

Finally,32{\displaystyle 3\Rightarrow 2} is trivial.{\displaystyle \square }

Counterexample

[edit]

The open mapping theorem may not hold for normed spaces that are not complete. A quickest way to see this is to note that theclosed graph theorem, a consequence of the open mapping theorem, fails without completeness. But here is a more concrete counterexample. Consider the spaceX ofsequencesx : N → R with only finitely many non-zero terms equipped with thesupremum norm. The mapT : X → X defined by

Tx=(x1,x22,x33,){\displaystyle Tx=\left(x_{1},{\frac {x_{2}}{2}},{\frac {x_{3}}{3}},\dots \right)}

is bounded, linear and invertible, butT−1 is unbounded. This does not contradict the bounded inverse theorem sinceX is notcomplete, and thus is not a Banach space. To see that it's not complete, consider the sequence of sequencesx(n) ∈ X given by

x(n)=(1,12,,1n,0,0,){\displaystyle x^{(n)}=\left(1,{\frac {1}{2}},\dots ,{\frac {1}{n}},0,0,\dots \right)}

converges asn → ∞ to the sequencex(∞) given by

x()=(1,12,,1n,),{\displaystyle x^{(\infty )}=\left(1,{\frac {1}{2}},\dots ,{\frac {1}{n}},\dots \right),}

which has all its terms non-zero, and so does not lie inX.

The completion ofX is thespacec0{\displaystyle c_{0}} of all sequences that converge to zero, which is a (closed) subspace of thep space(N), which is the space of all bounded sequences. However, in this case, the mapT is not onto, and thus not a bijection. To see this, one need simply note that the sequence

x=(1,12,13,),{\displaystyle x=\left(1,{\frac {1}{2}},{\frac {1}{3}},\dots \right),}

is an element ofc0{\displaystyle c_{0}}, but is not in the range ofT:c0c0{\displaystyle T:c_{0}\to c_{0}}. Same reasoning applies to showT{\displaystyle T} is also not onto inl{\displaystyle l^{\infty }}, for examplex=(1,1,1,){\displaystyle x=\left(1,1,1,\dots \right)} is not in the range ofT{\displaystyle T}.

Consequences

[edit]

The open mapping theorem has several important consequences:

The open mapping theorem does not imply that a continuous surjective linear operator admits a continuous linear section. What we have is:[9]

  • A surjective continuous linear operator between Banach spaces admits a continuous linear section if and only if the kernel is topologically complemented.

In particular, the above applies to an operator between Hilbert spaces or an operator with finite-dimensional kernel (by theHahn–Banach theorem). If one drops the requirement that a section be linear, a surjective continuous linear operator between Banach spaces admits a continuous section; this is theBartle–Graves theorem.[13][14]

Generalizations

[edit]

Local convexity ofX{\displaystyle X} orY{\displaystyle Y}  is not essential to the proof, but completeness is: the theorem remains true in the case whenX{\displaystyle X} andY{\displaystyle Y} areF-spaces. Furthermore, the theorem can be combined with the Baire category theorem in the following manner:

Open mapping theorem for continuous maps[12][15]LetA:XY{\displaystyle A:X\to Y} be acontinuous linear operator from a completepseudometrizable TVSX{\displaystyle X} onto a Hausdorff TVSY.{\displaystyle Y.}IfImA{\displaystyle \operatorname {Im} A} isnonmeager inY{\displaystyle Y} thenA:XY{\displaystyle A:X\to Y} is a (surjective) open map andY{\displaystyle Y} is a complete pseudometrizable TVS.Moreover, ifX{\displaystyle X} is assumed to be hausdorff (i.e. aF-space), thenY{\displaystyle Y} is also an F-space.

(The proof is essentially the same as the Banach or Fréchet cases; we modify the proof slightly to avoid the use of convexity,)

Furthermore, in this latter case ifN{\displaystyle N} is thekernel ofA,{\displaystyle A,} then there is a canonical factorization ofA{\displaystyle A} in the formXX/NαY{\displaystyle X\to X/N{\overset {\alpha }{\to }}Y}whereX/N{\displaystyle X/N} is thequotient space (also an F-space) ofX{\displaystyle X} by theclosed subspaceN.{\displaystyle N.}The quotient mappingXX/N{\displaystyle X\to X/N} is open, and the mappingα{\displaystyle \alpha } is anisomorphism oftopological vector spaces.[16]

An important special case of this theorem can also be stated as

Theorem[17]LetX{\displaystyle X} andY{\displaystyle Y} be twoF-spaces. Then every continuous linear map ofX{\displaystyle X} ontoY{\displaystyle Y} is aTVS homomorphism,where a linear mapu:XY{\displaystyle u:X\to Y} is a topological vector space (TVS) homomorphism if the induced mapu^:X/ker(u)Y{\displaystyle {\hat {u}}:X/\ker(u)\to Y} is a TVS-isomorphism onto its image.

On the other hand, a more general formulation, which implies the first, can be given:

Open mapping theorem[15]LetA:XY{\displaystyle A:X\to Y} be a surjectivelinear map from acompletepseudometrizable TVSX{\displaystyle X} onto a TVSY{\displaystyle Y} and suppose that at least one of the following two conditions is satisfied:

  1. Y{\displaystyle Y} is aBaire space, or
  2. X{\displaystyle X} islocally convex andY{\displaystyle Y} is abarrelled space,

IfA{\displaystyle A} is aclosed linear operator thenA{\displaystyle A} is an open mapping.IfA{\displaystyle A} is acontinuous linear operator andY{\displaystyle Y} is Hausdorff thenA{\displaystyle A} is (a closed linear operator and thus also) an open mapping.

Nearly/Almost open linear maps

A linear mapA:XY{\displaystyle A:X\to Y} between two topological vector spaces (TVSs) is called anearly open map (or sometimes, analmost open map) if for every neighborhoodU{\displaystyle U} of the origin in the domain, the closure of its imageclA(U){\displaystyle \operatorname {cl} A(U)} is a neighborhood of the origin inY.{\displaystyle Y.}[18] Many authors use a different definition of "nearly/almost open map" that requires that the closure ofA(U){\displaystyle A(U)} be a neighborhood of the origin inA(X){\displaystyle A(X)} rather than inY,{\displaystyle Y,}[18] but for surjective maps these definitions are equivalent.A bijective linear map is nearly open if and only if its inverse is continuous.[18]Every surjective linear map fromlocally convex TVS onto abarrelled TVS isnearly open.[19] The same is true of every surjective linear map from a TVS onto aBaire TVS.[19]

Open mapping theorem[20]If a closed surjective linear map from acompletepseudometrizable TVS onto a Hausdorff TVS isnearly open then it is open.

Theorem[21]IfA:XY{\displaystyle A:X\to Y} is a continuous linear bijection from acompletePseudometrizabletopological vector space (TVS) onto a Hausdorff TVS that is aBaire space, thenA:XY{\displaystyle A:X\to Y} is ahomeomorphism (and thus an isomorphism of TVSs).

Webbed spaces are a class oftopological vector spaces for which the open mapping theorem and theclosed graph theorem hold.

See also

[edit]

References

[edit]
  1. ^Trèves 2006, p. 166.
  2. ^Rudin 1973, Theorem 2.11.
  3. ^Vogt 2000, Theorem 1.6.
  4. ^Vogt 2000, Lemma 1.4.
  5. ^The first part of the proof ofRudin 1991, Theorem 2.11.
  6. ^abRudin 1991, Theorem 4.13.
  7. ^Vogt 2000, Lemma 1.5.
  8. ^Vogt 2000, Corollary 1.7.
  9. ^abTao, Terence (February 1, 2009)."245B, Notes 9: The Baire category theorem and its Banach space consequences".What's New.
  10. ^Rudin 1973, Corollary 2.12.
  11. ^Rudin 1973, Theorem 2.15.
  12. ^abRudin 1991, Theorem 2.11.
  13. ^Sarnowski, Jarek (October 31, 2020)."Can the inverse operator in Bartle-Graves theorem be linear?".MathOverflow.
  14. ^Borwein, J. M.; Dontchev, A. L. (2003). "On the Bartle–Graves theorem".Proceedings of the American Mathematical Society.131 (8):2553–2560.doi:10.1090/S0002-9939-03-07229-0.hdl:1959.13/940334.MR 1974655.
  15. ^abNarici & Beckenstein 2011, p. 468.
  16. ^Dieudonné 1970, 12.16.8.
  17. ^Trèves 2006, p. 170
  18. ^abcNarici & Beckenstein 2011, pp. 466.
  19. ^abNarici & Beckenstein 2011, pp. 467.
  20. ^Narici & Beckenstein 2011, pp. 466−468.
  21. ^Narici & Beckenstein 2011, p. 469.

Bibliography

[edit]

This article incorporates material from Proof of open mapping theorem onPlanetMath, which is licensed under theCreative Commons Attribution/Share-Alike License.

Further reading

[edit]
Spaces
Properties
Theorems
Operators
Algebras
Open problems
Applications
Advanced topics
Basic concepts
Main results
Maps
Types of sets
Set operations
Types of TVSs
Retrieved from "https://en.wikipedia.org/w/index.php?title=Open_mapping_theorem_(functional_analysis)&oldid=1286830665"
Category:
Hidden categories:

[8]ページ先頭

©2009-2025 Movatter.jp