3.7. Completeness#

The material in this section is primarily based on [2, 41].

Recall from Sequences that in a Cauchy sequence, the points become increasingly closure to each other. We showed that every convergent sequence is a Cauchy sequence. However not every Cauchy sequence is a convergent sequence in a general metric space. This is an important issue to be examined in this section. There are metric spaces where every Cauchy sequence is also convergent. Such metric spaces are known as complete metric spaces.

In the remainder of this section, (X,d) and (Y,ρ) shall denote metric spaces.

3.7.1. Complete Metric Spaces#

Definition 3.46 (Complete metric space)

A metric space (X,d) is called complete if all of its Cauchy sequences converge in the space.

In other words, X is complete if every Cauchy sequence {xn} of X converges to a point xX.

Example 3.19 (Complete metric spaces)

  1. Rn with the standard metric d(x,y)=|xy| is complete.

  2. The Euclidean space Rn, with the standard Euclidean metric (2 distance), is complete.

Example 3.20 (The metric space of bounded functions)

Let X be a non-empty set. Let B(X) be the set of bounded (total) functions on X.

for any f,gB(X), define D:B(X)×B(X)R as:

D(f,g)sup{|f(x)g(x)||xX}.

Since both f and g are bounded, hence D(f,g) is a real number. Thus domD=B(X)×B(X). We claim that D is a metric:

  1. D(f,g) is nonnegative as it is the supremum of nonnegative numbers.

  2. D(f,f)=sup{|f(x)f(x)|}=0. Also, if D(f,g)=0 then f(x)=g(x)xX. Thus, f=g.

  3. By definition: D(f,g)=D(g,f).

  4. For the triangle inequality, we proceed as follows:

    Let f,g,hB(X). For each xX, we have:

    |f(x)g(x)||f(x)h(x)|+|h(x)g(x)|D(f,h)+D(h,g).

    Now, taking the supremum on the L.H.S. over all xX, we get:

    D(f,g)D(f,h)+D(h,g).

Next, we establish that (B(X),D) is a complete metric space. Towards this, introduce a zero function z:XR as z(x)=0.

  1. Let {fn} be a Cauchy sequence of B(X).

  2. Thus, for every ϵ>0, there exists n0 such that D(fn,fm)<ϵ for all m,n>n0.

  3. In particular, note that, by definition, |fn(x)fm(x)|D(fn,fm) for every xX.

  4. Thus, for any xX, |fn(x)fm(x)|<ϵ for all m,n>n0.

  5. Thus, {fn(x)} is a Cauchy sequence of real numbers at each x=X.

  6. Since, R is complete, hence {fn(x)} converges at each x=X.

  7. We define a new function f:XR as f(x)=limfn(x) for each x=X.

  8. Since |fn(x)fm(x)|<ϵ for all m,n>n0, hence |fn(x)f(x)|ϵ for all n>n0 and all xX.

  9. Thus, |f(x)|ϵ+|fn(x)| for all xX.

  10. Thus, since fn is bounded, hence f is also bounded.

  11. Thus, fB(X).

  12. Finally, D(fn,f)=sup{|fn(x)f(x)|}.

  13. Since |fn(x)f(x)|ϵ for all xX, hence, taking the supremum on the L.H.S., D(fn,f)ϵ.

  14. Since D(fn,f)ϵ for all n>n0, hence limfn=f.

  15. Thus, every Cauchy sequence in B(X) converges in B(X).

  16. Thus, (B(X),D) is a complete metric space.

3.7.2. Closed Subsets#

Theorem 3.58

Let (X,d) be a complete metric space. Then a subset A of X is closed if and only if (A,d) is a complete metric space in its own right.

Proof. Let A be closed.

  1. Let {xn} be a Cauchy sequence of A.

  2. Then, {xn} is a Cauchy sequence of X (since AX).

  3. Thus, {xn} is convergent, since X is complete.

  4. Let limxn=x.

  5. Since A is closed and {xn} of A is convergent, hence xA due to Theorem 3.33.

  6. Thus, every Cauchy sequence of A converges in A.

  7. Thus, (A,d) is complete.

Let A be complete. We shall show that it contains all its closure points.

  1. Let xX be a closure point of A.

  2. By Theorem 3.31, there is a sequence {xn} of A that converges to x, i.e., limxn=x.

  3. Since {xn} converges in X, hence {xn} is a Cauchy sequence of X due to Proposition 3.19.

  4. But then, {xn} is a Cauchy sequence of A.

  5. Since, A is complete, hence {xn} converges in A.

  6. Thus, xA.

  7. Thus, A contains all its closure points. A is closed.

Recall that the diameter of a set is defined to be the supremum of distances between all pairs of points of the set.

Theorem 3.59

Let (X,d) be a complete metric space. Let {An} be a sequence of closed, nonempty subsets of X such that An+1An for each n and

limndiamAn=0.

Then the intersection n=1An consists of precisely one point.

This result is due to G. Cantor. A similar result was seen in nested interval property for real line.

Proof. Define:

A=n=1An.

We shall first prove that A cannot have more than one point.

  1. Assume x,yA.

  2. Thus, x,yAnn.

  3. Thus, 0d(x,y)diamAnn.

  4. But then, limndiamAn=0 implies that d(x,y)=0.

  5. But, d(x,y)=0x=y as d is a metric (Identity of indiscernibles).

  6. Thus, if x,yn=1An, then x=y.

  7. Thus, A contains at most one point.

We now show that A cannot be empty.

  1. For each n choose xnAn. It is possible due to axiom of choice.

  2. Since An+pAn, hence d(xn+p,xn)diamAn holds for every n,p.

  3. Thus, {xn} is a Cauchy sequence of X.

  4. Since X is complete, {xn} is convergent.

  5. Hence, the limit x=limnxn exists and xX.

  6. Since xmAn for all mn, hence x is a closure point of An for each n (Theorem 3.31).

  7. But since An is closed, hence xAn for each n.

  8. Thus, xA.

3.7.3. Nowhere Dense Sets#

Definition 3.47 (Nowhere dense/Rare)

A subset A of (X,d) is nowhere dense if its closure has an empty interior; i.e.,

intclA=.

It is also called a rare set.

Theorem 3.60

A is nowhere dense if and only if X(clA) is dense in X.

Proof. Let B=clA.

Recall from Theorem 3.15 that:

X(intB)=cl(XB)intB=X(cl(XB)).

Now,

intclA=intB=X(cl(XB))=cl(XB)=Xcl(X(clA))=XX(clA) is dense in X.

Example 3.21

The set of integers Z is nowhere dense in R.

Corollary 3.1

If a set is open and dense, then its complement is rare (nowhere dense).

Proof. Let A be open and dense.

  1. B=XA is closed.

  2. Thus, clB=B.

  3. Hence, X(clB)=XB=A.

  4. But A is dense, hence B must be nowhere dense due to Theorem 3.60.

Theorem 3.61

The complement of a rare (nowhere dense) set is dense.

Proof. Let AX be rare (nowhere dense).

  1. The interior of its closure is empty.

  2. Thus, its closure clA contains no open sets.

  3. Thus, A contains no open sets.

  4. Thus, every open set in X intersects with XA.

  5. Thus, XA is dense.

Proposition 3.20

The subset of a rare (nowhere dense) set is rare.

Proof. Let AB and B be nowhere dense. Now

ABclAclBintclAintclB.

Since B is nowhere dense, hence intclB is empty. This in turn implies that intclA is empty. Thus, A is nowhere dense.

Proposition 3.21

The boundary of a closed set is nowhere dense (rare).

Proof. Let A be closed. Thus, A=clA. Since bdA=clAintA, hence bdAA for a closed set.

  1. Assume bdA has a nonempty interior.

  2. Let x be an interior point of bdA.

  3. Since xbdAA, hence x is an interior point of A.

  4. But boundary of A doesn’t includes its interior points.

  5. Hence, a contradiction.

  6. Thus, bdA has an empty interior.

  7. Finally intclbdA = \interior \boundary A = \EmptySetsince\boundary A$ is closed.

  8. Thus, bdA is rare.

3.7.4. Cantor Set#

Definition 3.48 (Cantor set)

Cantor set C is a subset of [0,1]. It is constructed as follows:

  1. Let C0=[0,1].

  2. Trisect C0 into [0,13],(13,23),[23,1].

  3. Remove the middle open interval (13,23) and form C1=[0,13][23,1].

  4. Note that C1 is a disjoint union of 2=21 closed intervals.

  5. Trisect each closed interval of C1 and remove the middle open interval from each one of the trisections in an identical manner.

  6. Let C2=[0,19][29,13][23,79][89,1].

  7. Note that C2 is a union of 4=22 disjoint closed intervals each of length 19=132.

  8. Inductively build Cn+1 from Cn using this procedure.

  9. Note that Cn is a union of 2n disjoint closed intervals of length 13n each.

  10. When Cn+1 is constructed from Cn, we get 2n+1 disjoint intervals of length 13n+1 each.

  11. Clearly Cn+1Cn by construction for all n.

  12. Define the Cantor set as:

    Cn=1Cn.

We next discuss different properties of the Cantor set.

Property 3.1 (Characterization in ternary expansions)

Consider the set E[0,1] whose every element has a ternary (base 3) expansion of only 0s and 2s; i.e., for every xE, we have a representation:

x=.d1d2d3 where di{0,2}

Then, E=C. In other words:

C=E{n=1dn3n, where dn{0,2}}.

Proof. In the ternary (base 3), representation, each number x[0,1] can be written as:

x=.d1d2d3 where di{0,1,2}

such that

x=i=1di3i.

For example

1=.2222222...=2i=113i=21/32/3=212=1.

It is possible that a number has two different ternary expansions. We can see that 13Ck for every k. Hence, 13C.Interestingly,

13=.1=.022222....

has two different ternary representations. One is a finite representation, and the other is an infinite representation involving only 0s and 2s. We say that 13C as it has a ternary representation consisting of only 0s and 2s.

We have to prove two things:

  1. EC. Every number in [0,1] which has a ternary expansion containing only 0s and 2s belongs to C.

  2. CE. Every number in C has a ternary expansion containing only of 0s and 2s.

Note that if A=[a,a+13n] is a closed interval contained in Cn then:

  1. [a,a+13n+1] is the first closed interval drawn from A in Cn+1.

  2. [a+23n+1,a+13n] is the second closed interval drawn from A in Cn+1.

EC

  1. Let x=.d1d2 with di{0,2}.

  2. Start with A0=[0,1].

  3. Given A0:

    1. If d1=0, then let A1 be the first closed interval drawn from A0 : [0,13] contained in C1.

    2. Otherwise, if d1=2, then let A1 the second closed interval drawn from A0 : [23,1] contained in C1.

  4. Given An1=[a,a+13n1],

    1. If dn=0, then let An be the first closed interval drawn from An1 contained in Cn.

    2. Otherwise (if dn=2), then let An be the second closed interval drawn from An1 contained in Cn.

  5. Inductively, we can keep picking closed interval An contained in Cn for every digit in the ternary expansion of x.

  6. Thus, xC.

CE

  1. For every xC, we can construct a ternary expansion as follows.

  2. If x[0,13], then d1=0.

  3. Otherwise if x[23,1], then d1=2.

  4. Subsequently, whenever the first closed interval is chosen, then dn=0 and whenever the second closed interval is chosen, then dn=2.

  5. Thus, x has a ternary expansion consisting entirely of 0s and 2s.

  6. Thus, xE.

Property 3.2

Cantor set doesn’t contain any open interval.

Proof. Note that the total length of disjoint closed intervals in Cn is 2n3n=(23)n.

  1. Assume that there is an open interval (a,b)C.

  2. Then, the length of the interval is ba>0.

  3. But then there exists an n such that (23)n<ba.

  4. Then, for all kn, (a,b) cannot be contained in Ck.

  5. We arrive at a contradiction.

Property 3.3

Cantor set has an empty interior.

Proof. Assume C has a nonempty interior and xintC.

  1. Then there is a neighborhood (xϵ,x+ϵ)C.

  2. But, C doesn’t contain any open intervals.

  3. We arrive at a contradiction.

Thus, C has an empty interior.

Property 3.4

Cantor set is a closed nowhere dense subset of R.

Proof. C is an (infinite) intersection of closed sets. Hence C is closed. clC=C. C has an empty interior. Thus, C is nowhere dense.

Property 3.5

The total length of the removed intervals from [0,1] to get C equals 1.

Proof. At the n-th step, we remove 2n1 open intervals of length 3n each.

Thus, total length removed in n-th step is 12(23)n.

Thus, total length removed is:

n=112(23)n=122/312/3=1.

Property 3.6

Cantor set is uncountable. In particular:

cardC=card2N=cardR=c.

Recall that c denotes the cardinality of the continuum (Definition 1.92). The notation 2N was introduced in Theorem 1.34 to describe power sets. Theorem 1.33 established that

2NP(N)R.

Proof. Recall that two sets are called equivalent (AB) if there is a bijective mapping between them.

Recall from Property 3.1 that:

C={n=1dn3n, where dn{0,2}}.
  1. Thus, each xC can be identified with a sequence d:N{0,2} given by d={dn}.

  2. Thus, we have a bijective mapping between C and the set {0,2}N.

  3. Thus, C{0,2}N.

  4. This, in turn can be identified with a sequence c:N{0,1} where c={cn} with cn=0 if dn=0 and cn=1 if dn=2.

  5. This gives us a bijective mapping between {0,2}N and {0,1}N.

  6. Recall that 2N={0,1}N with 2={0,1} (Theorem 1.34).

  7. Thus C2N and

    cardC=card2N=cardR=c.

3.7.5. Meager Sets#

Definition 3.49 (Meager set)

A union of countably many rare (nowhere dense) sets is said to be of first category or a meager set.

In other words, a subset A is called meager (or of first category) if there exists a sequence {An} of nowhere dense subsets such that

A=n=1An.

Definition 3.50 (Co-meager set)

The complement of a meager set is called co-meager.

Definition 3.51 (Non-meager set)

A subset that is not meager is said to be non-meager or of second category.

Example 3.22

Consider the metric space R.

  1. Singleton sets are rare (nowhere dense) in R as they are closed and their interior is empty.

  2. The set of natural numbers N is rare since it is closed and its interior is empty (no open intervals in N).

  3. The set Q is not rare since its closure is entire R. It is meager since it is a countable union of rare singleton sets.

  4. The set of irrational numbers I is co-meager as I=RQ.

  5. The set R is non-meager.

Theorem 3.62

The subset of a meager set is meager.

Proof. Let A be a meager set and let BA. Then,

A=n=1An

where An are rare (nowhere dense).

But then

B=BA=Bn=1An=n=1AnB.

Since subset of a nowhere dense set is nowhere dense, hence Bn=AnB are nowhere dense. Hence

B=n=1Bn

is meager (as it is a countable union of nowhere dense sets).

Theorem 3.63

The union of countably many meager sets is meager.

Proof. Recall from Theorem 1.27 that countable union of countable sets is countable.

  1. Let {Mn} be a countable collection of meager sets.

  2. Then, each Mn is a countable union of rare sets.

  3. Write Mn as iAn,i.

  4. Thus,

    M=nMn=niAn,i=n,iAn,i.
  5. The family {An,i} is countable.

  6. Thus, M is a countable union of rare sets.

  7. Thus, M is meager.

Corollary 3.2

Since co-meager sets are complements of meager sets:

  1. Superset of a co-meager set is co-meager.

  2. Countable intersection of co-meager sets is co-meager.

3.7.6. Baire Category Theorem#

Theorem 3.64 (Baire category theorem)

A (nonempty) complete metric space is non-meager in itself.

In other words:

  • A (nonempty) complete metric space is not a countable union of rare sets.

Proof. Our proof strategy is following:

  1. For contradiction, we assume that X is meager with X=iAi.

  2. We form a Cauchy sequence from points xiAi.

  3. Since X is complete, we claim its convergence x=limxi.

  4. We further show that the limit point x cannot belong to any Ai.

Let us assume that X is a complete metric space which is meager in itself; i.e., there exists a countable collections of rare sets (rare in X) such that X is their union.

X=k=1Ak

where Ak are rare (nowhere dense); i.e. intclAk=.

Recall from Theorem 3.60 that Bk=XclAk are dense in X. Also, Bk are open since they are complement of closed sets. Also, recall from Theorem 3.23 that a dense set has a nonempty intersection with every nonempty open set.

  1. B1=XclA1 is a nonempty open set.

  2. Thus, there exists an interior point x1 in B1 with an open ball B(x1,ϵ1)B1 where we can make ϵ1<12.

  3. B2 is nonempty, open and dense. Thus, C2=B2B(x1,ϵ12) is nonempty and open.

  4. Thus, we can choose x2C2 such that xB(x2,ϵ2)C2 with ϵ2ϵ12.

  5. Proceeding inductively in this manner, we obtain a sequence {xn} such that

    xnB(xn,ϵn)BnB(xn1,ϵn12).
  6. We have AnB(xn,ϵn)=. ϵnϵn12 with ϵ1<12.

  7. By definition, xmB(xn,ϵn2) for every m>n.

  8. Thus, d(xm,xn)<ϵn2 for every m>n.

  9. Since ϵ1<12, hence ϵn<12n+1.

  10. Thus, d(xm,xn)0 as m,n.

  11. Thus, {xn} is a Cauchy sequence.

  12. Since X is complete, hence every Cauchy sequence is convergent.

  13. Thus there exists xX such that x=limxn. To which Ak does x belong then?

  14. Fix some nN. For every m>n we have:

    d(x,xn)d(x,xm)+d(xn,xm)<d(x,xm)+ϵn2.
  15. Thus taking the limit m leading to d(x,xm)0, we get

    d(x,xn)ϵn2<ϵn.
  16. Hence xB(xn,ϵn) for all nN.

  17. Since AnB(xn,ϵn)=, hence xAn for all n.

  18. But then xX since X=nAn. A contradiction.

  19. Hence, X must be non-meager.

Theorem 3.65

If (X,d) is complete and X=n=1An, then at least one An is non-rare; i.e., intclAn for some n.

Proof. If every An were rare, then X would be a meager set. But as per Baire category theorem, X must be non-meager since it is complete.

Hence, at least one An would be non-rare.

Proposition 3.22

The set of irrational numbers is non-meager.

Proof. Recall that R=QI where I is the set of irrational numbers.

  1. R is complete. Hence R is non-meager.

  2. Q is meager as it is a countable union of singletons which are rare sets.

  3. Countable union of meager sets is meager.

  4. Thus, if I was meager, then R would be meager which is not true.

  5. Hence, II must be non-meager.

Theorem 3.66 (Interior of a meager set)

A meager set has an empty interior in a complete metric space.

Proof. Let MX be meager. Then, we can write M as

M=n=1An

such that An are rare in X.

Let Bn=XclAn. Then Bn are dense, nonempty and open in X. Hence BnU is nonempty and open for every nonempty open set UX.

  1. Let U be an arbitrary nonempty open set in X.

  2. There exists x1B1U such that x1B(x1,ϵ1)B1U where ϵ1<12.

  3. B2 is dense and open so C2=B2B(x1,ϵ12) is nonempty and open.

  4. We can choose a point x2B(x2,ϵ2)C2 where ϵ2ϵ12.

  5. Proceeding in this manner, we choose points

    xnB(xn,ϵn)BnB(xn1,ϵn12)

    to form a sequence {xn}.

  6. The sequence {xn} is Cauchy, X is complete, hence x=limxn exists in X.

  7. Also, xB(xn,ϵn) for all n.

  8. In particular xB(x1,ϵ1)U and xB(xn,ϵn)Bn.

  9. Thus, xBnU for all n.

  10. Thus,

    xU(nBn)xU(n(XclAn))xU(X(n(clAn)))xU(X(nAn))xU(XM).
  11. Thus, U has a nonempty intersection with XM.

  12. Since U is arbitrary, hence XM intersects with every nonempty open set in X.

  13. Thus, XM is dense in X (Theorem 3.23).

  14. Thus, M has an empty interior (Theorem 3.24).

Corollary 3.3

If a set in a complete metric space has a non-empty interior, then it is not meager.

Example 3.23

Q in R

  1. The set Q is meager in R.

  2. R is complete.

  3. The interior of Q in R is empty.

  4. The closure of Q is R.

  5. Thus, while a meager set may have an empty interior, its closure need not have an empty interior. This is different from rare sets whose closure has an empty interior.

Q in Q

  1. Q by itself is not a complete metric space.

  2. Singletons are rare sets in Q.

  3. Thus, Q is meager in Q as it is a countable union of rare sets.

  4. However, the interior of Q is not empty. In fact it is whole of QQ.

  5. Theorem 3.66 does not apply since Q is not a complete metric space.

Theorem 3.67

The set [0,1] is uncountable.

Proof. We prove this using Baire category theorem.

  1. [0,1] is a complete metric space with the standard metric.

  2. Assume [0,1] to be countable.

  3. Then, [0,1]={xn}nN is an enumeration of [0,1].

  4. The singleton set {xn} is rare in [0,1].

  5. Then [0,1] being a countable union of rare sets would be meager.

  6. But Baire category theorem says that a complete metric space is non-meager.

  7. We have a contradiction.

  8. Thus, [0,1] must be uncountable.

3.7.7. Baire Spaces#

Definition 3.52 (Baire space)

A metric space is called a Baire space if every nonempty open set is not a meager set.

Theorem 3.68

Every complete metric space is a Baire space.

Proof. Let A be a nonempty open subset of a complete metric space (X,d). If it was meager, then it would have an empty interior (Theorem 3.66). Thus, it must be non-meager.

Hence X is a Baire space.

Theorem 3.69 (Characterization of Baire space)

For a metric space X, the following statements are equivalent:

  1. X is a Baire space.

  2. Every countable intersection of open dense sets is also dense.

  3. If X=n=1Fn and each Fn is a closed set, then the open set n=1intFn is dense.

Proof. (1) (2)

  1. Assume X is a Baire space.

  2. Let {An} be a sequence of open dense sets in X.

  3. Let A=n=1An. We need to show that A is dense.

  4. We will show that A has a nonempty intersection with every nonempty open set of X. Thus, claim that A is dense due to Theorem 3.23.

  5. Let OX be an arbitrary nonempty open set.

  6. Assume, for contradiction that AO=.

  7. Then, X=X=X(AO)=(XA)(XO).

  8. Thus,

    O=XO=((XA)(XO))O=(XA)O=(X(n=1An))O=n=1((XAn)O).
  9. Due to Corollary 3.1, XAn are rare.

  10. Due to Proposition 3.20, (XAn)O are rare.

  11. Thus, O being a countable union of rare sets, is meager.

  12. But, in a Baire space, every nonempty open set is not meager.

  13. We have a contradiction.

  14. Thus, AO must be nonempty.

  15. Thus, A has a nonempty intersection with every nonempty open set in A.

  16. Thus, A is dense in X.

(2) (3)

  1. We assume that every countable intersection of open dense sets is dense.

  2. Let {Fn} be a sequence of closed sets in X satisfying X=n=1Fn.

  3. Let An=intFn and let A=i=1An. By definition, A is open.

  4. Since Fn is closed, hence En=Fn(intFn) is its boundary.

  5. Due to Proposition 3.21, En is rare.

  6. Thus, the set E=n=1En is a meager set.

  7. Since En is closed and rare, hence XEn is open and dense (Theorem 3.61).

  8. By our hypothesis (2), the set

    XE=X(n=1En)=n=1(XEn)

    is also a dense set as it is a countable intersection of open dense sets XEn.

  9. Now, notice that:

    XA=X(i=1An)=n=1Fn(i=1(intFn))n=1[Fn(intFn)]=n=1En=E.
  10. And XAE implies XEA.

  11. Since XE is dense, hence A is also dense.

(3) (1). We need to show that every nonempty open set is non-meager.

  1. Let V be a nonempty open set. Assume V to be meager.

  2. Then V is a countable union of rare sets:

    V=n=1An

    where An are rare, thus intclAn=.

  3. We can write X as:

    X=(XV)V=(XV)A1A2=(XV)(clA1)(clA2).

    The last expression is correct since clAnX.

  4. In this form, X is a countable union of closed sets.

  5. By our hypothesis (3), the open set:

    (int(XV))(int(clA1))(int(clA2))=int(XV)

    is dense in X. Here, we used the fact that An are rare.

  6. Since int(XV)XV, hence XV is also dense in X.

  7. In particular V(XV) since a dense set has a nonempty intersection with every nonempty open set (Theorem 3.23).

  8. But this is impossible since V(XV)=.

  9. Thus, V cannot be not a meager set.

  10. We have established that any nonempty open V is not a meager set.

  11. Hence, X is a Baire space.

3.7.8. Completion#

Theorem 3.70

Let (X,d) be a metric space and let (Y,ρ) be a complete metric space. If f:XY with A=domf is a uniformly continuous function on A then f has a unique uniformly continuous extension to the closure of A.

Proof. If a sequence {xn} of A converges to a closure point x=limxn, then the sequence {f(xn)} of Y converges to some limit y=limf(xn).

  1. Let xclA.

  2. There exists a sequence {xn} of A such that limxn=x.

  3. Consider the sequence {f(xn)} of Y.

  4. Choose ϵ>0.

  5. Since f is uniformly continuous on A, hence there exists δ>0 such that ρ(f(x),f(y))<ϵ whenever d(x,y)<δ.

  6. Since {xn} is convergent and hence Cauchy, we can pick n0 such that d(xm,xn)<δ for all m,n>n0.

  7. Thus, ρ(f(xm),f(xn))<ϵ for all m,n>n0.

  8. Thus, {f(xn)} is a Cauchy sequence of Y.

  9. Since Y is complete, hence every sequence converges.

  10. Thus, there is a limit y=limf(xn).

For any sequence {xn} of A converging to x=limxn, the corresponding sequence {f(xn)} of Y has the same limit.

  1. Assume that x=limxn=limyn where {xn} and {yn} are two different sequences converging to x.

  2. Let u=limf(xn) and v=limf(yn). We claim that u=v.

  3. Consider the sequence {zn} defined as z2n=xn and z2n1=yn.

  4. It is easy to show that {zn} converges to x.

  5. Then, limf(zn) exists in Y.

  6. If a sequence converges, then all its subsequences converge to the same limit.

  7. Thus, limf(xn)=limf(yn)=limf(zn).

  8. Thus, u=v.

  9. Therefore limf(xn) is independent of the choice of sequence {xn} as long as x=limxn.

We define a function f:XY with domg=clA given the function f:XY with domf=A as:

f(x)=limf(xn)

where {xn} is any sequence converging to xclA; i.e., x=limxn.

f is well defined since limf(xn) is independent of the choice of the sequence {xn} converging to xclA.

We next establish that f is uniformly continuous.

  1. Let ϵ>0.

  2. We can choose δ>0 such that ρ(f(x),f(y))<ϵ whenever d(x,y)<δ for any x,yA.

  3. Now let, x,yclA satisfying d(x,y)<δ.

  4. Let {xn} and {yn} be convergent sequences of A with limxn=x and limyn=y.

  5. Then limd(xn,yn)=d(x,y) due to Theorem 3.34.

  6. Since d(x,y)<δ, we can pick n0 such that d(xn,yn)<δ for all n>n0.

  7. But since xn,ynA and f is uniformly continuous, we have, ρ(f(xn),f(yn))<ϵ for all n>n0.

  8. Since {f(xn)} and {f(yn)} are convergent sequences of Y, hence limρ(f(xn),f(yn))=ρ(f(x),f(y)) again due to Theorem 3.34.

  9. Thus, ρ(f(x),f(y))ϵ.

  10. Thus, f is uniformly continuous on clA.

Definition 3.53 (Completion of a metric space)

A complete metric space (Y,ρ) is called a completion of a metric space (X,d) if there exists an isometry f:(X,d)(Y,ρ) with domf=X such that f(X) is dense in Y.

If we think of X and f(X) as identical (up to an isometry), then we can think of X as a subset of Y.

Theorem 3.71

Any two completions of a metric space are isometric.

Proof. Let (Y1,ρ1) and (Y2,ρ2) be two different completions of (X,d).

  1. Then there are isometries f:XY1 and g:XY2 with domf=domg=X.

  2. g is an isometry with domg=X and rangeg=g(X).

  3. f1 is an isometry with domf1=f(X) and rangef1=X.

  4. Then, h=gf1 is an isometry from Y1 to Y2 with domh=f(X) and rangeh=g(X).

  5. f(X) is dense in Y1. Hence clf(X)=Y1.

  6. h is uniformly continuous (since it is an isometry).

  7. Y2 is complete.

  8. Then, due to Theorem 3.70, there exists a uniformly continuous extension h of h to all of Y1.

We have established that h is uniformly continuous.

We next show that h is an isometry.

  1. Let u,vY1 and z=h(u) and w=h(v).

  2. There is a sequence {xn} of X such that limf(xn)=u.

  3. There is a sequence {yn} of X such that limf(yn)=v.

  4. Let un=f(xn) and vn=g(yn).

  5. Let zn=g(un) and wn=g(yn).

  6. We have limzn=z and limwn=w.

  7. Since f is an isometry, hence d(xn,yn)=ρ1(un,vn).

  8. Since g is an isometry, hence d(xn,yn)=ρ2(zn,wn).

  9. Thus, ρ1(un,vn)=ρ2(zn,wn).

  10. Taking limits, we get ρ1(u,v)=ρ2(z,w).

  11. Thus, h is an isometry.

  12. Since it is an isometry, hence it is injective.

We next show that h is surjective (onto).

  1. Let vY2.

  2. There exists a sequence {xn} such that limg(xn)=v.

  3. There exists uY1 such that limf(xn)=u.

  4. Let f(xn)=un and g(xn)=vn.

  5. Then, h(un)=(gf1)(un)=g(xn)=vn.

  6. Thus, h(un)=vn.

  7. Then, since h is uniformly continuous, limun=u \implies limh(un)=limvn=v=h(u).

  8. Thus, for every vY2, there exists uY1 such that h(u)=v.

  9. Thus, h(u) is surjective.

Together since h is an isometric which is onto, hence Y1 and Y2 are isometric.

Remark 3.10

If f:(X,d)(Y,ρ) is an isometry with domf=X and Y is a complete metric space, then clf(X) is a completion of X.

Proof. Let Z = clf(X). Then Z is a closed subset of Y. By Theorem 3.58, (Z,ρ) is a complete subspace.

Now define g:(X,d)(Z,ρ) as:

g(x)=f(x).

Then, g is an isometry and g(X) is dense in Z. Thus, Z is a completion of X.

Theorem 3.72

Every metric space has a unique (up to an isometry) completion.

Proof. We prove this theorem by constructing a metric space which is a completion of a given metric space.

  1. Let (X,d) be a metric space.

  2. Fix an element aX.

  3. Now, for every xX, we introduce a function fx:XR defined as:

    fx(y)=d(x,y)d(y,a)yX.

    Note that domfx=X.

  4. From the triangular inequality we have:

    |fx(y)|=|d(x,y)d(y,a)|d(x,a)yX.
  5. Thus, for a given x and fixed a, the parametrized function fx is bounded by d(x,a).

  6. Thus, fxB(X), the space of bounded functions from X to R.

  7. We established in Example 3.20 that the metric space B(X) is a complete metric space.

We next show that the mapping f:XB(X) given by xfx is an isometry. Recall that the distance between two functions g,hB(X) is given by:

D(g,h)=sup{|g(x)h(x)||xX}.

Now for some x,zX with corresponding function fx,fzB(X), for any yX, we have:

|fx(y)fz(y)|=|d(x,y)d(y,a)[d(z,y)d(y,a)]|=|d(x,y)d(z,y)|d(x,z).

At the same time:

|fx(z)fz(z)|=|d(x,z)d(z,z)|=d(x,z).

Thus,

D(fx,fz)=sup{|fx(y)fz(y)||yX}=d(x,z).

Thus, f is an isometry. Since (B(X),D) is a complete metric space, hence as per Remark 3.10, (clf(X),D) is a completion of X.

We have already established in Theorem 3.71 that any two completions of X are isometric to each other. Thus, the completion is unique up to an isometry.

3.7.9. Isometries#

Theorem 3.73 (Isometries preserve closed sets)

Let (X,d) and (Y,ρ) be complete metric spaces. Let f:(X,d)(Y,ρ) be an isometry. Let CX be a closed set in X. Then, f(C) is closed.

Proof. 1. Let {yn} be a convergent sequence of f(C).

  1. Let y=limyn.

  2. Recall that an isometric is injective.

  3. Thus, yn=f(xn) for some xnC for every n.

  4. Since {yn} is convergent, hence {yn} is Cauchy.

  5. Now, for any m,nN

    d(xm,xn)=ρ(ym,yn)

    as f is isometric.

  6. Thus, {xn} is a Cauchy sequence of C.

  7. Since X is complete, hence every Cauchy sequence converges.

  8. Let x=limxn.

  9. Since C is closed, hence xC.

  10. By continuity of f, we have f(x)=y.

  11. Thus, y=f(x)f(C).

  12. Thus, every convergent sequence of C converges in C.

  13. Thus, f(C) is closed.