Kreyszig 2.5 Compactness and Finite Dimension

Problem 1. Show that Rn\mathbb{R}^n and Cn\mathbb{C}^n are not compact.

Solution

To show that Rn\mathbb{R}^n and Cn\mathbb{C}^n are not compact, we can utilize the Heine-Borel theorem, which characterizes compact subsets of Rn\mathbb{R}^n. The theorem states that a subset of Rn\mathbb{R}^n is compact if and only if it is closed and bounded.

For Rn\mathbb{R}^n:

  1. Closedness: By definition, Rn\mathbb{R}^n is closed because it contains all its limit points; every convergent sequence in Rn\mathbb{R}^n has a limit that is also in Rn\mathbb{R}^n.

  2. Boundedness: However, Rn\mathbb{R}^n is not bounded. To see this, consider the sequence {(xk)}k=1\{(x_k)\}_{k=1}^{\infty} where xk=(k,0,0,,0)x_k = (k, 0, 0, \ldots, 0) in Rn\mathbb{R}^n. This sequence has no upper bound within Rn\mathbb{R}^n because for any given MRM \in \mathbb{R}, there exists an NN such that for all n>Nn > N, xn>Mx_n > M.

Thus, since Rn\mathbb{R}^n is not bounded, it cannot be compact according to the Heine-Borel theorem.

For Cn\mathbb{C}^n:

  1. Closedness: Cn\mathbb{C}^n is also closed because it includes all its limit points; it is the entire space of nn-tuples of complex numbers.

  2. Boundedness: We can show that Cn\mathbb{C}^n is not bounded in a similar manner to Rn\mathbb{R}^n. Consider the sequence of complex numbers {(zk)}k=1\{(z_k)\}_{k=1}^{\infty} where zk=(k,0,,0)z_k = (k, 0, \ldots, 0) in Cn\mathbb{C}^n, where kk represents the complex number k+0ik + 0i. This sequence, too, has no bound within Cn\mathbb{C}^n for the same reasons as in Rn\mathbb{R}^n.

Thus, Cn\mathbb{C}^n is not bounded and, hence, not compact.

Note that Cn\mathbb{C}^n is isomorphic to R2n\mathbb{R}^{2n} since each complex number corresponds to a pair of real numbers (the real and imaginary parts). Therefore, the non-compactness of Cn\mathbb{C}^n follows from the non-compactness of R2n\mathbb{R}^{2n}.

In conclusion, neither Rn\mathbb{R}^n nor Cn\mathbb{C}^n is compact because, although both are closed, neither is bounded.

\blacksquare

The detailed explanation of why the sequence {(xk)}k=1\{(x_k)\}_{k=1}^{\infty} with xk=(k,0,0,,0)x_k = (k, 0, 0, \ldots, 0) in Rn\mathbb{R}^n has no upper bound is as follows:

For any chosen real number MM, no matter how large, there exists a natural number NN such that for all k>Nk > N, the value of kk is greater than MM. This means that the first component of the vector xkx_k is greater than MM. The formal expression of this statement is:

MR,NN:k>N,xk>M \forall M \in \mathbb{R}, \exists N \in \mathbb{N} : \forall k > N, x_k > M

This expression states that for any real number MM one might consider as a potential upper bound, there exists a point in the sequence, beyond the NN th term, where the elements of the sequence exceed MM. Thus, for any MM in R\mathbb{R}, we can find elements in the sequence {(xk)}\{(x_k)\} that are larger than MM, demonstrating that the sequence does not have an upper bound in Rn\mathbb{R}^n.

The absence of an upper bound implies that the sequence is unbounded. According to the Heine-Borel theorem, a necessary condition for a subset of Rn\mathbb{R}^n to be compact is that it must be bounded. Since the sequence {(xk)}\{(x_k)\} is unbounded in Rn\mathbb{R}^n, it follows that Rn\mathbb{R}^n itself is not compact as it fails to satisfy the boundedness condition required by the theorem.


Problem 2. Show that a discrete metric space XX consisting of infinitely many points is not compact.

Solution

To prove that an infinite discrete metric space XX is not compact, we use the definition of compactness in metric spaces. A metric space is compact if every open cover has a finite subcover.

In a discrete metric space, the metric is defined such that d(x,x)=0d(x, x) = 0 for all xXx \in X, and d(x,y)=1d(x, y) = 1 for all xyx \neq y. This implies that each point in XX is isolated from every other point. We can then consider an open cover of XX consisting of the open balls {B(x,12)}\{B(x, \frac{1}{2})\} for each xXx \in X. Each of these balls is indeed an open set because it contains no points other than its center.

Since XX contains infinitely many points, the collection {B(x,12)}\{B(x, \frac{1}{2})\} is an infinite cover for XX. If XX were compact, there would exist a finite subcover that still covers XX. However, this is not possible because each open ball in our cover contains exactly one point of XX and no two balls contain the same point. Thus, no finite collection of these balls can cover the entirety of XX.

Consequently, there is no finite subcover possible for the cover {B(x,12)}\{B(x, \frac{1}{2})\}, which means that the discrete metric space XX cannot be compact.

\blacksquare


Problem 3. Give examples of compact and noncompact curves in the plane R2\mathbb{R}^2.

Solution

Compact Curves:

  1. Unit Circle: The set of all points (x,y)(x, y) such that x2+y2=1x^2 + y^2 = 1. This curve is closed and bounded.

  2. Square: The boundary defined by the set of all points (x,y)(x, y) with vertices at (±1,±1)(\pm1, \pm1). It is a closed and bounded shape.

  3. Triangle: The boundary formed by connecting points (0,0)(0, 0), (1,0)(1, 0), and (0,1)(0, 1). Each edge is a closed line segment, making the whole triangle compact.

  4. Closed Disk: The set of all points (x,y)(x, y) satisfying x2+y2r2x^2 + y^2 \leq r^2 for a fixed rr. This includes all points within and on the boundary, constituting a closed and bounded set.

Noncompact Curves:

  1. Ray: The set of points (x,y)(x, y) forming a ray extending from the origin indefinitely, like {(t,t)t0}\{(t, t) | t \geq 0\}. This curve is unbounded.

  2. Hyperbola: The set of points (x,y)(x, y) satisfying xy=1xy = 1. This curve extends to infinity in all directions.

  3. Infinite Line: A line like y=xy = x that extends without bound in both directions.

  4. Logarithmic Spiral: Defined by the polar equation r=eθr = e^{\theta}, this curve winds away from the origin infinitely.

These examples illustrate the distinction in R2\mathbb{R}^2 between compact sets, which are both closed and bounded, and noncompact sets, which are not closed, not bounded, or both.


Problem 4. Show that for an infinite subset MM in the space ss to be compact, it is necessary that there are numbers γ1,γ2,\gamma_1, \gamma_2, \ldots such that for all x=(ξk(x))Mx = (\xi_k(x)) \in M we have ξk(x)γk|\xi_k(x)| \leq \gamma_k. (It can be shown that the condition is also sufficient for the compactness of MM.)

Solution

To show the necessity of the condition for compactness in the space ss, as defined in the problem statement, we need to demonstrate that for any sequence in a compact subset MM of ss, the elements of the sequence must be uniformly bounded by some sequence {γk}\{\gamma_k\}.

Consider the space ss as defined in the second image, where the metric dd is given by:

d(x,y)=j=112jξjηj1+ξjηj d(x, y) = \sum_{j=1}^{\infty} \frac{1}{2^j} \frac{| \xi_j - \eta_j |}{1 + | \xi_j - \eta_j |}

with x=(ξk)x = (\xi_k) and y=(ηk)y = (\eta_k) being elements of ss.

The metric dd is designed such that the "distance" it measures is the sum of a series of terms, each of which is a fraction of the absolute difference between the components of two elements xx and yy, scaled by 1/2j1/2^j. This series converges because each term is less than or equal to 1/2j1/2^j, and 1/2j\sum 1/2^j is a convergent geometric series.

Now let's consider the subset MsM \subset s. If MM is compact, then from the definition of compactness in metric spaces, every sequence in MM has a convergent subsequence. For a sequence {x(n)}\{x^{(n)}\} with x(n)=(ξk(n))x^{(n)} = (\xi_k^{(n)}) in MM, its convergence in ss means that for every ϵ>0\epsilon > 0, there exists an NN such that for all m,n>Nm, n > N, d(x(m),x(n))<ϵd(x^{(m)}, x^{(n)}) < \epsilon.

For compactness, we require that this sequence has a convergent subsequence in ss. Due to the definition of the metric, this means that for each jj, the sequence {ξj(n)}\{\xi_j^{(n)}\} must be Cauchy, and hence bounded. Therefore, there must exist a bound γj\gamma_j for each jj such that ξj(n)γj|\xi_j^{(n)}| \leq \gamma_j for all nn.

To see why the sequence {ξj(n)}\{\xi_j^{(n)}\} must be bounded, suppose it were not. If for some jj, {ξj(n)}\{\xi_j^{(n)}\} were unbounded, then we could choose ϵ\epsilon small enough (specifically ϵ<1/2j\epsilon < 1/2^j) and a subsequence {x(nk)}\{x^{(n_k)}\} such that ξj(nk)ξj(nk+1)>1|\xi_j^{(n_k)} - \xi_j^{(n_{k+1})}| > 1 for all kk, which would imply d(x(nk),x(nk+1))d(x^{(n_k)}, x^{(n_{k+1})}) would not converge to 0, contradicting the compactness of MM.

Therefore, for MM to be compact, it is necessary that there exist numbers γ1,γ2,\gamma_1, \gamma_2, \ldots such that for all x=(ξk(x))Mx = (\xi_k(x)) \in M we have ξk(x)γk|\xi_k(x)| \leq \gamma_k. This condition is known to be sufficient as well for the compactness of MM in the space ss, as a uniformly bounded and equicontinuous sequence in ss will have a convergent subsequence by the Arzelà-Ascoli theorem.

\blacksquare


Problem 5. A metric space XX is said to be locally compact if every point of XX has a compact neighborhood. Show that R\mathbb{R} and C\mathbb{C}, and more generally, Rn\mathbb{R}^n and Cn\mathbb{C}^n are locally compact.

Solution

To prove that R\mathbb{R}, C\mathbb{C}, and by extension Rn\mathbb{R}^n and Cn\mathbb{C}^n, are locally compact, we utilize the following concepts:

  1. A space is locally compact if each point has a compact neighborhood.

  2. A set is compact if every open cover has a finite subcover, which, in a metric space, translates to the set being closed and bounded, as per the Heine-Borel theorem.

  3. A neighborhood of a point includes an open set containing that point.

Detailed Proofs of Local Compactness

Detailed Proof for R\mathbb{R}:

For any point xRx \in \mathbb{R}, we can identify a neighborhood around xx, such as the open interval (xϵ,x+ϵ)(x - \epsilon, x + \epsilon) for some ϵ>0\epsilon > 0. The closure of this interval is the closed interval [xϵ,x+ϵ][x - \epsilon, x + \epsilon], which encompasses its limit points and is delimited by the points xϵx - \epsilon and x+ϵx + \epsilon. By the Heine-Borel theorem, as [xϵ,x+ϵ][x - \epsilon, x + \epsilon] is both closed and bounded within R\mathbb{R}, it is compact. Therefore, every point xx possesses a compact neighborhood in R\mathbb{R}, affirming its local compactness.

Detailed Proof for C\mathbb{C}:

Upon recognizing C\mathbb{C} as topologically equivalent to R2\mathbb{R}^2, for any zCz \in \mathbb{C}, we consider the open disk centered at zz, denoted D(z,ϵ)D(z, \epsilon) for some ϵ>0\epsilon > 0. This disk serves as a neighborhood of zz. The closure of D(z,ϵ)D(z, \epsilon), which consists of all points inside and on the boundary of the disk, constitutes a closed set. It is also bounded by the circumference of the disk. Thus, by the Heine-Borel theorem, the closure of D(z,ϵ)D(z, \epsilon) is compact in C\mathbb{C}, corroborating its local compactness.

Detailed Proof for Rn\mathbb{R}^n:

For an arbitrary point xRnx \in \mathbb{R}^n, we select the open ball B(x,ϵ)B(x, \epsilon) centered at xx with a radius ϵ>0\epsilon > 0. The closure of this ball, B(x,ϵ)\overline{B(x, \epsilon)}, which includes all points within and on the periphery of the sphere, is closed. Moreover, it is bounded as all points lie within a maximum distance ϵ\epsilon from xx. Consequently, B(x,ϵ)\overline{B(x, \epsilon)} is compact as per the Heine-Borel theorem, demonstrating that Rn\mathbb{R}^n is locally compact since xx has a compact neighborhood.

Detailed Proof for Cn\mathbb{C}^n:

Given that Cn\mathbb{C}^n aligns with R2n\mathbb{R}^{2n} topologically, each complex coordinate having a real and imaginary part, for any point zCnz \in \mathbb{C}^n, an open ball in R2n\mathbb{R}^{2n} can be centered at the point corresponding to zz with a radius ϵ>0\epsilon > 0. The closure of this ball is also a closed and bounded set in R2n\mathbb{R}^{2n}, and hence compact. This provides every point in Cn\mathbb{C}^n with a compact neighborhood, certifying local compactness.

Each proof underlines the principle that local compactness is evidenced by the ability to encase any point within a closed and bounded (thus compact) subset, meeting the local compactness criterion.


Problem 6. Show that a compact metric space XX is locally compact.

Proof

Let XX be a compact metric space. We aim to prove that for every point xx in XX, there exists a compact neighborhood around xx. In metric spaces, we have the luxury of using open balls as basic neighborhoods. For an arbitrary xXx \in X and for any positive real number ϵ\epsilon, the open ball B(x,ϵ)B(x, \epsilon) is an open set containing xx.

Due to the compactness of XX, any open cover has a finite subcover. Consider the collection of open balls {B(x,1n)}nN\{B(x, \frac{1}{n})\}_{n \in \mathbb{N}}, which is indeed an open cover of XX. By the compactness of XX, there exists a finite subcover of this collection, implying the existence of some NNN \in \mathbb{N} such that B(x,1N)B(x, \frac{1}{N}) is contained within an open set that is part of the finite subcover of XX.

The closure of B(x,1N)B(x, \frac{1}{N}), denoted by B(x,1N)\overline{B(x, \frac{1}{N})}, is a closed subset of the compact space XX. By the properties of compact spaces, closed subsets of compact spaces are also compact. Thus, B(x,1N)\overline{B(x, \frac{1}{N})} is compact and contains the open ball B(x,1N)B(x, \frac{1}{N}), which is a neighborhood of xx. This establishes that xx has a compact neighborhood.

Since the choice of xx in XX was arbitrary, and we have demonstrated that each point has a compact neighborhood, it follows that the metric space XX is locally compact.

This detailed proof leverages the Heine-Borel theorem and the properties of open and closed sets in metric spaces to demonstrate the local compactness of a compact metric space.

\blacksquare


Problem 7. If dimY<\dim Y < \infty in Riesz's lemma 2.5-4, show that one can even choose θ=1\theta = 1.

Proof Using Riesz's Lemma

Let us consider Riesz's lemma in the context where YY is a finite-dimensional subspace of ZZ, a subspace of a normed space XX. Riesz's lemma asserts that given a closed subspace YY which is a proper subset of ZZ, for every real number θ\theta in the interval (0,1), there exists a zZz \in Z such that z=1\|z\| = 1 and zyθ\|z - y\| \geq \theta for all yYy \in Y.

Suppose vZYv \in Z \setminus Y and denote the distance from vv to YY by aa, where a=inf{vy:yY}a = \inf\{\|v - y\| : y \in Y\}. Since YY is closed and finite-dimensional, it is also a known fact that closed balls in YY are compact. Thus, the infimum aa is actually achieved by some y0Yy_0 \in Y. We have vy0=a\|v - y_0\| = a and a>0a > 0 because vv is not in YY.

We proceed to define zz as the normalization of vy0v - y_0, so z=c(vy0)z = c(v - y_0) where c=1vy0=1ac = \frac{1}{\|v - y_0\|} = \frac{1}{a}. This normalization ensures that z=1\|z\| = 1.

For any yYy \in Y, we can express yy as y1=y0+c1yy_1 = y_0 + c^{-1}y, with y1y_1 also in YY due to the vector space properties of YY. The norm zy\|z - y\| is then c(vy0)y=cvy1\|c(v - y_0) - y\| = c\|v - y_1\|. Given that vv is closest to y0y_0 by the very definition of aa, it follows that cvy1cvy0=ca=1c\|v - y_1\| \geq c\|v - y_0\| = c \cdot a = 1. Consequently, zy1\|z - y\| \geq 1 for all yYy \in Y.

Since the choice of yy was arbitrary, this implies that zyθ\|z - y\| \geq \theta for any θ1\theta \leq 1. Thus, when YY is finite-dimensional, it is permissible to select θ=1\theta = 1 in Riesz's lemma. The lemma is thereby applicable for θ=1\theta = 1, which is due to the structure of the normed space and the finite-dimensionality of YY, guaranteeing the existence of such a zz with the specified characteristics.

\blacksquare

A little different approach

To show that θ=1\theta=1 can be chosen in Riesz's lemma under the condition that the dimension of YY is finite, we will analyze the proof of Riesz's lemma and demonstrate that if YY has a finite dimension, then the distance from any vZYv \in Z \setminus Y to YY can be made equal to 1, which implies that θ\theta can be taken as 1.

Riesz's Lemma states that for any two subspaces YY and ZZ of a normed space XX, with YY being closed and a proper subset of ZZ, for every θ\theta in the interval (0,1), there exists a zZz \in Z such that z=1\|z\| = 1 and zyθ\|z - y\| \geq \theta for all yYy \in Y.

Proof Using Riesz's Lemma

Suppose YY is a finite-dimensional subspace of ZZ. By the properties of finite-dimensional normed spaces, we know that closed balls in YY are compact. Let vZYv \in Z \setminus Y and denote its distance from YY by aa, that is, a=inf{vy:yY}a = \inf\{\|v - y\| : y \in Y\}. Since YY is closed and vv is not in YY, it follows that a>0a > 0.

In the finite-dimensional subspace YY, due to compactness, the infimum aa is actually attained for some y0Yy_0 \in Y. That is, there exists a y0Yy_0 \in Y such that vy0=a\|v - y_0\| = a. Now, define zz as a scaled vector of vy0v - y_0, specifically z=c(vy0)z = c(v - y_0), where c=1vy0=1ac = \frac{1}{\|v - y_0\|} = \frac{1}{a}. This scaling ensures that z=1\|z\| = 1.

Now, consider any yYy \in Y. We examine the distance from zz to yy. Note that any yy can be written as y1=y0+c1yy_1 = y_0 + c^{-1}y, where y1Yy_1 \in Y due to YY being a vector space and thus closed under addition and scalar multiplication. We calculate:

zy=c(vy0)y=cvy0c1y=cvy1. \|z - y\| = \|c(v - y_0) - y\| = c\|v - y_0 - c^{-1}y\| = c\|v - y_1\|.

Because vv is closer to y0y_0 than any other point in YY by the definition of y0y_0, it follows that cvy1cvy0=ca=1c\|v - y_1\| \geq c\|v - y_0\| = c \cdot a = 1. Therefore, for all yYy \in Y, zy1\|z - y\| \geq 1, which by the choice of our zz implies zyθ\|z - y\| \geq \theta for any θ1\theta \leq 1. Hence, in the case where YY has finite dimension, we can choose θ=1\theta = 1 in Riesz's lemma.

This shows that the lemma is not only true for any θ\theta in the open interval (0,1) but can be strengthened to include θ=1\theta = 1 when the subspace YY is of finite dimension. The lemma holds trivially for θ=1\theta = 1 because the normed space structure and finite dimensionality ensure the existence of such zz with the required properties.

\blacksquare


Problem 8. In Problem 7, Section 2.4, show directly (without using 2.4-5) that there is an a>0a > 0 such that ax2xa\|x\|_2 \leq \|x\|. (Use 2.5-7.)

Show directly that there is a constant a>0a > 0 such that ax2xa\|x\|_2 \leq \|x\| for a normed finite-dimensional vector space XX without using the theorem on equivalent norms.

Proof

Let XX be a finite-dimensional vector space equipped with two norms \|\cdot\| and 2\|\cdot\|_2, where 2\|\cdot\|_2 is the standard Euclidean norm. Consider the unit sphere SS in XX with respect to 2\|\cdot\|_2, that is, S={xX:x2=1}S = \{x \in X : \|x\|_2 = 1\}.

Since XX is finite-dimensional, SS is compact with respect to 2\|\cdot\|_2. Now, define a mapping T:SRT: S \to \mathbb{R} by T(x)=xT(x) = \|x\| for all xSx \in S. This mapping is continuous because the norms are continuous functions, and by the Corollary 2.5-7, since SS is compact, TT attains its maximum and minimum values on SS.

Let m=min{T(x):xS}m = \min \{T(x) : x \in S\}. Since all norms on a finite-dimensional space are positive definite, we have m>0m > 0 because if m=0m = 0, there would exist an xSx \in S such that x=0\|x\| = 0, which implies x=0x = 0, contradicting the fact that xx is on the unit sphere SS.

Now, for any xXx \in X with x0x \neq 0, we can write xx as x=x2(xx2)x = \|x\|_2 \cdot \left(\frac{x}{\|x\|_2}\right). Notice that xx2S\frac{x}{\|x\|_2} \in S, hence xx2m\left\|\frac{x}{\|x\|_2}\right\| \geq m. Multiplying both sides by x2\|x\|_2, we get x=x2xx2mx2\|x\| = \|x\|_2 \cdot \left\|\frac{x}{\|x\|_2}\right\| \geq m \|x\|_2.

Set a=ma = m, which is the positive minimum value of TT on the compact set SS. We have established that ax2xa\|x\|_2 \leq \|x\| for all xXx \in X, where a>0a > 0.

This completes the proof, establishing the existence of a positive constant aa that provides a lower bound for the ratio of the norms \|\cdot\| and 2\|\cdot\|_2 on a finite-dimensional vector space XX.

\blacksquare


Problem.9 If XX is a compact metric space and MXM \subseteq X is closed, show that MM is compact.

Proof

Consider XX, a metric space endowed with a metric dd, and let MXM \subseteq X be a closed subset. Our objective is to substantiate the compactness of MM predicated on the compactness of the ambient space XX.

Compactness in a metric space is defined such that a subset MM of XX is compact if every open cover of MM admits a finite subcover. Let us take an arbitrary open cover O\mathcal{O} of MM, constituted by a family of open sets in XX such that every point in MM resides within some member of O\mathcal{O}.

Given the closure of MM in XX, its complement XMX \setminus M is open in XX. Enhance the open cover O\mathcal{O} of MM by annexing the open set XMX \setminus M, thereby generating a new open cover O\mathcal{O}' that extends over the entirety of XX, for it encompasses every point in XX.

The compact nature of XX necessitates that the open cover O\mathcal{O}' of XX must possess a finite subcover, designated as O\mathcal{O}''. This finite subcover aptly covers all points in XX, and by extension, all points in MM.

From the finite subcover O\mathcal{O}'', excise the set XMX \setminus M should it be included. The residual compendium of sets within O\mathcal{O}'' thus forms a finite subcollection originating from the initial cover O\mathcal{O}, which adequately covers MM. Consequently, MM is furnished with a finite subcover from its open cover O\mathcal{O}.

Ergo, MM aligns with the compactness criterion. We have thus rigorously delineated, utilizing the axioms of metric topology alongside the attributes of closed sets nestled within compact spaces, that a closed subset MM of a compact metric space XX is necessarily compact.

This consummates the proof, and we have methodically demonstrated, consistent with the tenets of metric space theory and the inherent properties of closed subsets within compact spaces, that a closed subset MM of a compact metric space XX must itself exhibit compactness.

\blacksquare