M7180 Funkcionální analýza II M7180 Functional analysis II Michal Veselý Text vychází ze zdroj˚u [1–4] Masarykova univerzita, Pˇrírodovˇedecká fakulta Ústav matematiky a statistiky Se sazbou v systému LATEX pomáhal Jakub Juránek Tato publikace vznikla v rámci projektu Fond rozvoje Masarykovy univerzity (projekt MUNI/FR/1133/2018) realizovaného v období 1/2019–12/2019. Contents 0 Linear operators 1 0.1 Preliminaries and examples . . . . . . . . . . . . . . . . . . . . . . . 1 0.2 Continuity and boundedness . . . . . . . . . . . . . . . . . . . . . . 2 0.3 Inverse operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 0.4 Adjoint operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 0.5 Spectrum of operator . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1 Completely continuous operators 18 1.1 Preliminaries and examples . . . . . . . . . . . . . . . . . . . . . . . 18 1.2 Basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 1.3 Self-adjoint operator in Hilbert space . . . . . . . . . . . . . . . . . . 28 2 Derivative in Banach spaces 32 2.1 Weak and strong derivative . . . . . . . . . . . . . . . . . . . . . . . 32 2.2 Convex function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.3 Tangent functional . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 3 Strictly and uniformly convex spaces 40 3.1 Strictly convex space . . . . . . . . . . . . . . . . . . . . . . . . . . 40 3.2 Uniformly convex space . . . . . . . . . . . . . . . . . . . . . . . . 42 3.3 Projection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 3.4 Smooth space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 4 Fixed point theorems 49 4.1 Topological degree . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 4.2 Brouwer and 1. and 2. Schauder theorem . . . . . . . . . . . . . . . 50 5 Integration in Banach spaces 53 5.1 Preliminaries and basic definitions . . . . . . . . . . . . . . . . . . . 53 5.2 Bochner integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 5.3 Gelfand, Pettis, and Darboux integral . . . . . . . . . . . . . . . . . . 57 0. Linear operators 0.1 Preliminaries and examples Let X,Y be linear spaces. Definition 0.1. Any map F : X → Y satisfying F(λx+ µy) = λF(x)+ µF(y) for all x,y ∈ X and scalars λ,µ is called a linear operator. Remark 0.2. For operator F, we can write Fx instead of F(x). Example 0.3. The operator I : X → X given by x → x is called the identical operator. Obviously, I is linear. If X is normed, then I is continuous (by the Heine definition of continuity: xn → x =⇒ Fxn → Fx). Example 0.4. Let H be a Hilbert space and H1 its (closed) subspace. We know that H = H1 ⊕ H⊥ 1 , i.e., all x ∈ H can be uniquely expressed as x = x1 + x2, where x1 ∈ H1, x2 ∈ H⊥ 1 . We put Px = x1 for x ∈ H. The operator P is called the (orthogonal) projection. Evidently, it is linear and continuous. Example 0.5. Let us consider the linear space C[a,b] and define the operator T : C[a,b] → C[a,b] by the formula T f(t) = b a k(t,s)f(s)ds, f ∈ C[a,b], t ∈ [a,b], with the so-called core k ∈ C([a,b]×[a,b]). We can see that T is linear. If we consider the norm f = max t∈[a,b] |f(t)|, f ∈ C[a,b], (0.1) i.e., the norm of the uniform convergence, then T is continuous. Example 0.6. Let us consider the linear space C1 [a,b] of functions with continuous derivatives on [a,b] and define the operator D: C1 [a,b] → C[a,b], Df(t) = f (t), f ∈ C1 [a,b], t ∈ [a,b]. This operator is called the differential operator and it is linear. In spacesC1 [a,b],C[a,b], let us consider the norm from (0.1). We prove that D is not continuous: the sequence of fn(t) = sin(nt) n , n ∈ N, satisfies fn → 0 as n → ∞, but the sequence of Dfn(t) = fn(t) = cos(nt) does not converge to 0. 1 0.2 Continuity and boundedness Throughout this section, let X = (X, · ) = (X, · X ),Y = (Y, · ) = (Y, · Y ) be normed linear spaces. Definition 0.7. An operator L: X → Y is called bounded if it maps any bounded set into a bounded set. Theorem 0.8. Any continuous linear operator L: X → Y is bounded. Proof. We assume the opposite. Let the operator L not be bounded. Then, there exists a sequence {xn}∞ n=1 ⊆ X such that xn X ≤ c for all n ∈ N and some c ∈ R and Lxn Y > n, n ∈ N. (0.2) Since L is continuous, there exists δ ∈ (0,1) such that, for all x ∈ X satisfying x X < δ, we have Lx Y ≤ 1. Let us choose n0 ∈ N such that c < δn0. The inequality xn X n < δ is valid for all n ≥ n0. Hence, we get Lxn Y ≤ n for all n ≥ n0, which is a contradiction with (0.2). Theorem 0.9. Any bounded linear operator L: X → Y is continuous at 0 ∈ X. Proof. Let us assume the opposite. Let the operator L not be continuous at 0 ∈ X. Then, there exist ε > 0 and a sequence {xn}∞ n=1 ⊆ X such that xn X < 1 n , Lxn Y ≥ ε, n ∈ N. Let us denote yn = nxn for n ∈ N. The sequence {yn}∞ n=1 ⊆ X is bounded, but the sequence {Lyn}∞ n=1 ⊆ Y is not bounded, because Lyn Y ≥ nε, n ∈ N, which is a con- tradiction. Theorem 0.10. If a linear operator L: X → Y is continuous at x0 ∈ X, then L is continuous at any point (vector) of X. Proof. For all ε > 0, there exists δ > 0 such that, for x ∈ X satisfying x−x0 < δ, we have Lx−Lx0 < ε. Let x1 ∈ X be arbitrarily given and y ∈ X satisfy y−x1 < δ. Then, (y−x1 +x0)−x0 < δ, and thus L(y−x1 +x0)−Lx0 < ε, i.e., Ly−Lx1 < ε, which proves that L is continuous at x1. 2 Remark 0.11. The definition of bounded linear operators can be reformulated as follows. A linear operator L: X → Y is called bounded if there exists c > 0 such that Lx Y ≤ c x X , x ∈ X. Definition 0.12. Let L: X → Y be a bounded linear operator. The number inf{c ∈ R; Lx Y ≤ c x X for x ∈ X} is denoted by L and called the norm of L. Theorem 0.13. Let L: X → Y be a bounded linear operator. Then, L = sup x X ≤1 Lx Y = sup x=0 Lx Y x X . Proof. We denote λ = sup{ Lx Y ; x ∈ X, x X ≤ 1}. Firstly, we can see that sup x X ≤1 Lx Y = sup x=0 Lx Y x X , where it suffices to consider the linearity of L and x X = 1. Therefore, Lx Y x X ≤ λ, x = 0, i.e., Lx Y ≤ λ x X for all x ∈ X. Considering Definition 0.12, we have L ≤ λ. Let ε > 0 be arbitrary. Then, there exists xε ∈ X such that xε = 0 and λ −ε ≤ Lxε Y xε X . However, from Definition 0.12, we obtain Lx Y ≤ L · x X for all x ∈ X. Thus, λ −ε ≤ L . The arbitrariness of ε gives λ = L . Remark 0.14. The set of all continuous linear operators L: X → Y is denoted by L (X,Y). For L1,L2 ∈ L (X,Y) and a scalar k, we put (L1 +L2)(x) = L1x+L2x, x ∈ X, (kL1)(x) = kL1(x), x ∈ X. Evidently, L (X,Y) forms a linear space. This space is normed (with respect to the norm of operators introduced above). Indeed, for the triangular inequality, it is sufficient to consider (L1 +L2)x Y ≤ L1x Y + L2x Y ≤ L1 + L2 for all L1,L2 ∈ L (X,Y) and x X ≤ 1, x ∈ X. Moreover, from Definition 0.12, we have the inequality Lx Y ≤ L · x X , x ∈ X, L ∈ L (X,Y). In the case when X = Y, we can write only L (X) (instead of L (X,X)). 3 Remark 0.15. Recall that norms − 1 and − 2 on X are called equivalent if there exist α,β > 0 such that α x 1 ≤ x 2 ≤ β x 1, x ∈ X. Two norm are equivalent if and only if the topologies generated by them are same. For any space with a finite dimension, all norms are equivalent. Theorem 0.16. Let X have a finite dimension. Any linear operator L: X → Y is con- tinuous. Proof. According to Remark 0.15, we can choose a norm of X. Let e1,...,en be a base of X. For x ∈ X,x = λ1e1 +···+λnen, we put x X = |λ1|+···+|λn|. It is seen that Lx Y = λ1Le1 +···+λnLen Y ≤ max{ Le1 Y ,..., Len Y }· x X . Remark 0.17. A basic example of continuous linear operators is given by functionals f on X with the norm f = sup x ≤1 |f(x)|. For details, we refer to the course Functional Analysis I. We add: 1. If X has a finite dimension, then f is realized. In addition, the norm of any continuous linear functional is realized on the closed unit ball with the center 0 if and only if the Banach space X is reflexive (the so-called James characteristic). 2. If any linear functional on X is bounded, then the dimension of X is finite. 3. The so-called Bishop–Phelps theorem says that the set of all continuous linear functionals on a Banach space X, whose norms are realized on the closed unit ball with the center 0, is a dense subset of the dual space X . Example 0.18. We mention a series of examples. a) Let us consider the space C[−1,1] with the norm f = max t∈[−1,1] |f(t)| and the functional L f = 9 f(−1)−2f(0)+ f 1 4 . If f ≤ 1, then |L f| ≤ 9 f +2 f + f ≤ 12. Especially, L ≤ 12. On the contrary, let us consider a function g ∈C[−1,1] for which max t∈[−1,1] |g(t)| = 1 and g(−1) = 1, g(0) = −1, g 1 4 = 1. We get Lg = 12. Therefore, L = 12. 4 b) For the space from a), we consider the functional L f = 1 −1 sgn(t)f(t)dt. Since |L f| ≤ 1 −1 |f(t)|dt ≤ f 1 −1 dt = 2 f , we have L ≤ 2. Let choose ε ∈ (0,1) and put gε (t) = t/ε for t ∈ (−ε,ε) and gε (t) = sgn t for others t. Then, gε = 1 and |Lgε | = 2−ε. Now, we see that L = 2. c) For the space l2 , we consider L{xn}∞ n=1 = x1 +x2. For x = {xn}∞ n=1 ∈ l2 satisfying x = |x1|2 +|x2|2 +··· ≤ 1, we get |Lx|2 ≤ (|x1|+|x2|)2 ≤ 2 |x1|2 +|x2|2 ≤ 2 |x1|2 +|x2|2 +··· ≤ 2. Therefore, L ≤ √ 2. For x = 1 √ 2 , 1 √ 2 ,0,0, ... , we have x = 1 and |Lx| = √ 2. Thus, L = √ 2. We also know (from the Riesz theorem) that there exists h ∈ l2 such that Lx = x,h , x ∈ l2 . We see that h = {1,1,0,0, ...}. In addition, we know that L = h , which gives L = √ 1+1 = √ 2. Our knowledge of the dual space gives a powerful tool to compute the norm of L. d) We compute the norm of the functional L: {xn}∞ n=1 → ∞ ∑ n=1 xn n in l1 and l2 . For x = {xn}∞ n=1 ∈ l1 , we have |Lx| = ∞ ∑ n=1 xn n ≤ ∞ ∑ n=1 |xn| = x . Therefore, L ≤ 1. For x = {1,0,0, ...}, we have x = 1, |Lx| = 1, consequently, L = 1. 5 We can also use our knowledge of the dual space. The dual space of l1 is l∞ . There exists just one {an}∞ n=1 ∈ l∞ satisfying Lx = ∞ ∑ n=1 anxn, x = {xn}∞ n=1 ∈ l1 , where L = {an}∞ n=1 l∞ . Obviously, an = 1/n, n ∈ N. We get L = 1 n ∞ n=1 l∞ = 1. Now, we consider L in l2 . Since {1/n}∞ n=1 ∈ l2 , for x = {xn}∞ n=1 ∈ l2 , we have |Lx| = ∞ ∑ n=1 xn n ≤ ∞ ∑ n=1 1 n2 · ∞ ∑ n=1 |xn|2 = π √ 6 · x . Thus, L ≤ π/ √ 6. For y = √ 6/(πn) ∞ n=1 ∈ l2 , we have y = 1 and |Ly| = √ 6 π ∞ ∑ n=1 1 n2 = π √ 6 , which yields that L = π/ √ 6. e) Let us consider X = C1 [a,b], Y = C[a,b], and the operator (see Example 0.6) D: X → Y, Df(t) = f (t), f ∈ C1 [a,b], t ∈ [a,b]. We consider the standard norm (see (0.1)) in Y and the norm f C1 = max t∈[a,b] |f(t)|+ max t∈[a,b] f (t) in X. Since Df = max t∈[a,b] |f (t)| ≤ max t∈[a,b] |f(t)|+ max t∈[a,b] |f (t)| = f C1 , f ∈ C1 [a,b], the operator D is continuous and D ≤ 1. To prove D = 1, it is sufficient to consider fn(t) = sin(nt) n , n ∈ N, for which fn C1 = 1+ 1 n , Dfn = 1, where n is sufficiently large. f) Let us consider the operators L1,L2 ∈ L l2 given by L1x = 0,x1, x2 2 , x3 3 , ... , x = {xn}∞ n=1 ∈ l2 , L2x = {0,x1,x2,x3, ...}, x = {xn}∞ n=1 ∈ l2 . Since the operator L2 is isometric, we see that L2 = 1. For L1, we consider arbitrary x = {xn}∞ n=1 ∈ l2 . We have L1x 2 = ∞ ∑ n=1 xn n 2 ≤ ∞ ∑ n=1 |xn|2 = x 2 , which gives L1 ≤ 1. For y = {1,0,0, ...}, we have y = 1 and L1y = 1. Altogether, L1 = 1. 6 0.3 Inverse operator In this section, let X,Y be linear spaces. Definition 0.19. Let L : X → Y be an arbitrary operator. We put R(L) = {y ∈ Y; there exists x ∈ X : Lx = y}. We say that the operator L has an inverse if, for all y ∈ R(L), there exists just one x ∈ X such that Lx = y. In this case, the map R(L) → X given by y → x is called the inverse operator of L and it is denoted by L−1 . Theorem 0.20. Let L : X → Y be linear and have an inverse. Then, L−1 : Y → X is linear as well. Proof. Note that the range R(L) of the operator L, i.e., the domain D L−1 of the inverse operator L−1 , is a linear space. Let y1,y2 ∈ R(L). It suffices to prove the identity L−1 (α1y1 +α2y2) = α1L−1 y1 +α2L−1 y2 (0.3) for all scalars α1,α2. Put Lx1 = y1, Lx2 = y2. We know that L(α1x1 +α2x2) = α1y1 +α2y2. (0.4) According to Definition 0.19, we see that L−1 y1 = x1 and L−1 y2 = x2. Thus, we have α1L−1 y1 +α2L−1 y2 = α1x1 +α2x2. At the same time, from Definition 0.19 and (0.4), we get α1x1 +α2x2 = L−1 (α1y1 +α2y2), which gives (0.3). Lemma 0.21. Let M be a dense subset of a Banach space Y. Any y = 0, y ∈ Y, can be expressed in the form y = ∞ ∑ n=1 yn, i.e., y = lim n→∞ (y1 +y2 +···+yn), where yn ∈ M and yn ≤ 3 y 2n , n ∈ N. Proof. At first, we choose y1 ∈ M in such a way that the inequality y−y1 ≤ y 2 is valid. Next, we choose y2 ∈ M so that y−y1 −y2 ≤ y 4 . In general, we choose yn ∈ M so that y−y1 −···−yn ≤ y 2n , n ∈ N. 7 Such a choice is possible, because M is dense in Y. Now, y− m ∑ n=1 yn → 0 as m → ∞, i.e., y = ∞ ∑ n=1 yn, and we have y1 = y1 −y+y ≤ y1 −y + y ≤ 3 y 2 , y2 = y2 +y1 −y+y−y1 ≤ y−y1 −y2 + y−y1 ≤ 3 y 4 , ... yn = yn +yn−1 +···+y1 −y+y−y1 −···−yn−1 ≤ y−y1 −···−yn + y−y1 −···−yn−1 ≤ y 2n + y 2n−1 = 3 y 2n . Theorem 0.22 (Banach). Let X,Y be Banach spaces and let L : X → Y be a bounded, bijective, and linear operator. Then, the inverse operator L−1 is bounded as well. Proof. In Y, we consider the sets Mk ⊆ Y, k ∈ N, of all elements y ∈ Y for which L−1 y ≤ k y . All element y ∈ Y belongs to some Mk. Thus, Y = ∞ k=1 Mk. According to the Baire theorem, at least one of Mk, say Mn, is dense in some ball B. Inside B, we consider the set P of all elements z such that β < z − y0 < α, where 0 < β < α, y0 ∈ Mn. We move the set P so that the center is 0, i.e., let us consider the set P0 = {z ∈ Y; β < z < α}. We prove that some MN is dense in P0. If z ∈ P∩Mn, then z−y0 ∈ P0 and L−1 (z−y0) ≤ L−1 z + L−1 y0 ≤ n( z + y0 ) ≤ n( z−y0 +2 y0 ) = n z−y0 1+ 2 y0 z−y0 ≤ n z−y0 1+ 2 y0 β . (0.5) The term n 1+ 2 y0 β 8 does not depend on z. So, we put N = 1+ n 1+ 2 y0 β . From (0.5), we have z − y0 ∈ MN. Since Mn is dense in P, we have that MN is dense in P0. Let y = 0, y ∈ Y, be arbitrary. We can choose such a number λ that β < λy < α, i.e., λy ∈ P0. Since MN is dense in P0, we can construct a sequence of yk ∈ MN which converges to λy. Then, yk/λ converges to y. It is obvious that {xk/µ}∞ k=1 ⊆ MN for arbitrary µ = 0 if {xk}∞ k=1 ⊆ MN. Therefore, the set MN is dense in Y {0} and, consequently, it is dense in Y. We consider arbitrary y = 0, y ∈ Y. According to Lemma 0.21, we can expand y into the series y = y1 +y2 +··· , where yk ∈ MN and yk ≤ 3 y 2k , k ∈ N. In X, we consider the series L−1 y1 +L−1 y2 +···, where we put xk = L−1 yk. This series converges to some x ∈ X, because xk = L−1 yk ≤ N yk ≤ N 3 y 2k , k ∈ N, x ≤ ∞ ∑ k=1 xk ≤ 3N y ∞ ∑ k=1 1 2k = 3N y . Since the series of xk is convergent and the operator L is continuous, we have Lx = Lx1 +Lx2 +··· = y1 +y2 +··· = y. Therefore, x = L−1 y. We also know that L−1 y = x ≤ 3N y , where N does not depend on y. This estimation is valid for arbitrary y = 0. Thus, the operator L−1 is bounded. Remark 0.23. Let X,Y be Banach spaces. The symbol ˜L (X,Y) denotes the set of all bijective, continuous, and linear operators X → Y. Theorem 0.24. Let X,Y be Banach spaces. Let L0 ∈ ˜L (X,Y) and L ∈ L (X,Y), where L < 1 L−1 0 . Then, the bounded operator (L0 +L)−1 exists on Y, i.e., L1 = L0 +L ∈ ˜L (X,Y). Proof. We choose y ∈Y and consider the map B: X → X, Bx = L−1 0 y−L−1 0 (Lx). From L < L−1 0 −1 , it follows that B is a contraction. Indeed, Bx1 −Bx2 ≤ L−1 0 · L · x1 −x2 , x1,x2 ∈ X. 9 Since X is complete, according to the Banach theorem, there exists just one x ∈ X such that x = Bx = L−1 0 y−L−1 0 (Lx), i.e., L1x = L0x+Lx = y. If L1 ¯x = y for some ¯x ∈ X, then ¯x is also a fixed point of B, and therefore ¯x = x. Thus, for all y ∈ Y, there exists only one solution of the equation L1x = y in X, i.e., for L1, there exists the inverse operator L−1 1 on Y. Considering Theorem 0.22, L−1 1 is bounded. Remark 0.25. According to the previous theorem, ˜L (X,Y) forms an open subset in L (X,Y), where X,Y are Banach spaces. Theorem 0.26 (Neumann). Let X be a Banach space, let I be the identical operator on X, and let L: X → X be a bounded linear operator, where L < 1. Then, the operator (I −L)−1 exists on X, it is bounded, and it can be expressed in the form (I −L)−1 = ∞ ∑ k=0 Lk , where Lk = L◦L◦···◦L k . Proof. The existence on X and the boundedness of (I −L)−1 come from Theorem 0.24 (also from treatments below). Because of L < 1, we have ∞ ∑ k=0 Lk ≤ ∞ ∑ k=0 L k < ∞. (0.6) The space X is complete. Thus, considering (0.6), the treated infinite sum of Lk is a bounded linear operator. For arbitrary n ∈ N∪{0}, we have (I −L) n ∑ k=0 Lk = n ∑ k=0 Lk (I −L) = I −Ln+1 . Taking into account to Ln+1 ≤ L n+1 → 0 as n → ∞, we get (I −L) ∞ ∑ k=0 Lk = ∞ ∑ k=0 Lk (I −L) = I, which yields (I −L)−1 = ∞ ∑ k=0 Lk . 10 0.4 Adjoint operator Throughout this section, let X,Y be normed linear spaces. Definition 0.27. Let L ∈ L (X,Y) and let g ∈ Y , i.e., let g be a continuous linear functional on Y. Let us consider the continuous linear functional f = g◦L ∈ X and the map g ∈ Y → f ∈ X . This map L : Y → X is called the adjoint operator of L. Remark 0.28. We also denote f(x) = f,x . Thus, we can write f,x = g,Lx , i.e., L g,x = g,Lx . Remark 0.29. Let L,L1,L2 ∈ L (X,Y) and let k be a scalar. Immediately, from Definition 0.27, we see: 1. L is linear; 2. (L1 +L2) = L1 +L2; 3. (kL) = kL ; 4. L is continuous. Example 0.30. Let us consider a linear (continuous) operator L: Rn → Rm given by a matrix (lij). The map y = Lx can be expressed as the system yi = n ∑ j=1 lijxj, i ∈ {1,...,m}, and any functional f : Rn → R as f(x) = n ∑ j=1 fjxj, where fj = f(ej) for the standard base e1,...,en of Rn . From f(x) = g(Lx) = m ∑ i=1 giyi = m ∑ i=1 n ∑ j=1 gilijxj = n ∑ j=1 xj m ∑ i=1 gilij, we obtain fj = f(ej) = m ∑ i=1 gilij, j ∈ {1,...,n}. Since f = L g, the operator L is given by the transpose matrix. Theorem 0.31. If L ∈ L (X,Y), then L = L . Proof. Obviously, it holds L g,x = | g,Lx | ≤ g · Lx ≤ g · L · x for all x ∈ X and g ∈ Y . Thus, L g ≤ g · L , i.e., L ≤ L . 11 Now, we prove the opposite inequality. Let x0 ∈ X, Lx0 = 0. We put y0 = Lx0 Lx0 ∈ Y. It is seen that y0 = 1. Due to a well-known corollary of the Hahn–Banach theorem, there exists a functional g such that g = 1 and g,y0 = 1, i.e., g,Lx0 = Lx0 . From Lx0 = g,Lx0 = | L g,x0 | ≤ L g · x0 ≤ L · g · x0 = L · x0 , we get that L ≤ L . Let H be a Hilbert space and let L : H → H be a bounded linear operator. We know that there exists a map τ which assigns to any element y ∈ H the continuous linear functional (τy)(x) = x,y ∈ H . Moreover, this map is an isometry. For the operator L , we consider the map ˜L = τ−1 L τ, which is a bounded linear operator on H. One can easily show that Lx,y = x, ˜L y , x,y ∈ H. Since L = L and the maps τ and τ−1 are isometries, we have the identity ˜L = L . Definition 0.32. In a Hilbert space H, the above mentioned operator ˜L : H → H is called the adjoint operator of L: H → H. Remark 0.33. It should be emphasized that Definition 0.32 differs from Definition 0.27. For a general Banach space X and a bounded linear operator L: X → X, the adjoint operator of L is defined on X . The operator ˜L is sometimes called the Hermitian adjoint. We write only L (instead of ˜L ) and speak about the adjoint operator of L. It should be remembered that, in Hilbert spaces, the concept of adjoint operators differs from the one in general Banach spaces. For H, it is seen that the adjoint operator of a bounded linear operator L: H → H can be defined as the operator L : H → H which satisfies Lx,y = x,L y , x,y ∈ H. Definition 0.34. Let H be a Hilbert space. A bounded linear operator L: H → H is called self-adjoint if Lx,y = x,Ly , x,y ∈ H. Definition 0.35. Let H be a Hilbert space and let L: H → H be a linear operator. A (closed) subspace H1 of H is called invariant with respect to L: H → H if x ∈ H1 implies Lx ∈ H1. Remark 0.36. Let H be a Hilbert space and let L: H → H be a bounded linear operator. If H1 is a (closed) subspace of H, which is invariant with respect to L, then its orthogonal complement H⊥ 1 is invariant with respect to L . Indeed, if y ∈ H⊥ 1 , then x,L y = Lx,y = 0, x ∈ H1, because Lx ∈ H1. Especially, if L is self-adjoint, then the orthogonal complement of any invariant subspace is invariant with respect to L as well. 12 0.5 Spectrum of operator Let L: Cn → Cn be a linear operator. A number λ ∈ C is called an eigenvalue of L if Lx = λx for some non-zero x ∈ Cn . Any such solution x is called an eigenvector of L. The set of all eigenvalues is called the spectrum of the operator L and all other values λ are called regular; i.e., λ is a regular value if the operator L − λI has an inverse. In this case, the operator (L−λI)−1 is defined on the entire space Cn and (as well as any linear operator on a space with a finite dimension) it is bounded (continuous). In the space with a finite dimension, there are two possibilities: 1. the equation Lx = λx has a non-zero solution, i.e., λ is an eigenvalue of L—the operator (L−λI)−1 does not exist; 2. the bounded operator (L−λI)−1 is defined in the whole space, i.e., λ is a regular value. If the operator L is defined on a space whose dimension is infinite, then we have the third possibility: 3. the operator (L−λI)−1 exists, i.e., the equation Lx = λx has only the zero solution, but this operator is not defined on the whole space (and it is not necessarily bounded). Let X be a complex Banach space. Definition 0.37. A number λ ∈ C is called a regular value of a bounded linear operator L: X → X if the operator Rλ = (L−λI)−1 , called the resolvent of L, is defined in the whole space X. The set of all non-regular values is called the spectrum of L and it is denoted by σ(L). The spectrum includes all eigenvalues of the operator L. Indeed, if (L−λI)x = 0 for some x = 0, then (L−λI)−1 does not exist. The set of all eigenvalues of L is called the point (or discrete) spectrum and the corresponding x = 0 are called eigenvectors. The remaining part of the spectrum, i.e., the set of all λ, for which the inverse operator (L−λI)−1 exists, but is not defined in the whole space X, is called the continuous spectrum. Theorem 0.38. The set of all regular values of a bounded linear operator L: X → X is open, i.e., the spectrum is a closed set. Proof. Let λ be a regular value of L. Then, L−λI ∈ ˜L (X) = ˜L (X,X). Let δ ∈ C satisfy |δ| < 1 ||(L−λI)−1 || . From Theorem 0.24, we have L − (λ +δ)I ∈ ˜L (X). Hence, λ + δ is a regular value of L and the set of all regular values is open. Theorem 0.39. Let L: X → X be a bounded linear operator and let |λ| > L . Then, λ is a regular value of L. 13 Proof. Because of L < |λ|, applying Theorem 0.26, we know that the operator Rλ = (L−λI)−1 = − 1 λ I − L λ −1 exists on X and it is bounded. Therefore, λ is a regular value. Remark 0.40. Theorem 0.39 can be specified as follows. Let r = lim n→∞ n Ln , where it is possible to show that this limit exists for any bounded linear operator L: X → X. The spectrum of L is in the closed circle with the radius r and the centre in 0. The value r is called the spectral radius of L and r ≤ L . For example, the operator L: C[0,1] → C[0,1] (we consider the norm from (0.1)) given by L: f(t) → t 1 0 f(x)dx, t ∈ [0,1], f ∈ C[0,1], has the norm L = 1 and r = lim n→∞ n ||Ln|| = 1 2 . We add that, for any bounded linear operator L: X → X, it holds r = sup{|λ|; λ ∈ σ(L)}. Example 0.41. We define the operator T : C[0,1] → C[0,1] by the formula T : f(t) → f t2 , t ∈ [0,1], f ∈ C[0,1], where we consider the norm from (0.1). At first, we determine T , which is easy. We have T = sup{ T f ; f ≤ 1} ≤ 1. For f0 ≡ 1, we obtain f0 = 1 and also T f0 = 1. Thus, T = 1. We find eigenvalues of T. We find such λ that the equation T f = λ f has non-zero solutions. We know that σ(T) ⊆ {z ∈ C; |z| ≤ T = 1}. For λ = 0 and T f = λ f = 0, we have f ≡ 0 and, therefore, 0 is not an eigenvalue. For λ = 1 and T f = λ f = f, all constant functions are solutions of the equation f t2 = f(t), t ∈ [0,1]. It remains to investigate such λ that |λ| ∈ (0,1],λ = 1. From T f(t) = f t2 = λ f(t), t ∈ [0,1], we obtain f(t) = 1 λ f t2 = 1 λ2 f t4 = ··· = 1 λn f t2n 14 for all t ∈ [0,1] and n ∈ N. We see that f(t) = λn f t2−n . If |λ| < 1, then f ≡ 0, because λn → 0 as n → ∞ and f is bounded on [0,1]. Let us consider the last case when |λ| = 1 and λ = 1. In this case, we get f(0) = λ f(0), f(1) = λ f(1) and, consequently, f(0) = f(1) = 0. Let t ∈ (0,1). From f(t) = 1 λn f t2n and t2n → 0 as n → ∞, it follows that f ≡ 0. Therefore, the set of all eigenvalues is {1}. It remains to find out other values belonging to the spectrum of T. Thus, we analyse solutions of the equation T f −λ f = g in C[0,1]. We know, when the operator T −λI is not injective. So, we are interested in the case, when it is not surjective. For t ∈ [0,1], we obtain f t2 = g(t)+λ f(t) and f(t) = g t1/2 +λ f t1/2 = g t1/2 +λg t1/4 +λ2 f t1/4 . By induction, one can obtain f(t) = n−1 ∑ j=0 λ j g t2−1−j +λn f t2−n , t ∈ [0,1]. Because of the boundedness of g, for |λ| < 1, the series above converges and λn f t2−n → 0 as n → ∞ for all f. Therefore, the continuous function f(t) = ∞ ∑ j=0 λ j g t2−1−j , t ∈ [0,1], is a solution of the equation T f −λ f = g. If |λ| = 1,λ = 1, then the equation T f −λ f = g has no solution for some g ∈C[0,1]. For example, for λ = −1, there exists a continuous function g ∈ C[0,1] such that g t2−1−j = (−1)j j for given t ∈ (0,1), j ∈ N. Altogether, we have σ(T) = {λ ∈ C; |λ| = 1}. 15 Example 0.42. Now, on the space C[0,1] with the norm from (0.1), we consider the operator ˜T : f(t) → t 0 f(x)dx, t ∈ [0,1], f ∈ C[0,1]. Its norm is ˜T = 1. Let us identify the point spectrum of the operator ˜T. We know that λ is an eigenvalue if there exists a non-zero solution of the equation ˜T f = λ f. So, we are looking for a non-trivial continuous function f ∈ C[0,1] with the property that t 0 f(x)dx = λ f(t), t ∈ [0,1]. We see that λ f(0) = 0 and that the function f has to have a continuous derivative if λ = 0. If λ = 0, then t 0 f(x)dx = 0, t ∈ [0,1], i.e., f ≡ 0. Therefore, 0 is not an eigenvalue of ˜T. If λ = 0, from the mentioned identity, we obtain f(t) = λ f (t), t ∈ [0,1]. Solutions of this equation are functions f(t) = Ket/λ . Since f(0) = 0, we get K = 0. Thus, the operator ˜T has no eigenvalues. Now, we determine the whole spectrum of ˜T. We need to find the values λ for which the considered operator is surjective, i.e., we need to determine when the equation ˜T f − λ f = g has solutions for all g ∈ C[0,1]. Let λ = 0. Since ˜T f(0) = 0, for a function g such that g(0) = 0, any solution does not exist. Hence, 0 ∈ σ( ˜T). In the case when λ = 0, we are looking for solutions of the equation t 0 f(x)dx−λ f(t) = g(t), t ∈ [0,1], where g ∈ C[0,1] is a given function. Let h be the appropriate primitive function of f. The aim is to solve the differential equation h−λh = g. Of course, this equation has solutions. Thus, we get σ ˜T = {0}. Example 0.43. On the space l∞ , we consider the operator R: {x1,x2,x3, ...} → {x2,x3, ...}. It is easy to verify that R ∈ L (l∞ ) and that R = 1. We know that σ(R) ⊆ {λ ∈ C; |λ| ≤ 1}. If |λ| ≤ 1, the equation Rx = λx has the non-trivial solution xλ = 1,λ,λ2 , ... . 16 Since xλ ∈ l∞ for |λ| ≤ 1, we have σ(R) = {λ ∈ C; |λ| ≤ 1}. Let us consider the same operator R, but on l1 . Of course, R ∈ L l1 and R = 1. The equation Rx = λx has (again) the non-trivial solution xλ . But, for |λ| = 1, this element is not in l1 . However, xλ ∈ l1 for all λ satisfying |λ| < 1. Therefore, we get that all such λ are in the point spectrum. Since the spectrum σ(R) is a closed set which contains the point spectrum, we have again σ(R) = {λ ∈ C; |λ| ≤ 1}. 17 1. Completely continuous operators 1.1 Preliminaries and examples Throughout this section, let X,Y be Banach spaces. Definition 1.1. An operator L: X → Y is called completely continuous if it maps any bounded set into a precompact set. Remark 1.2. If X has a finite dimension, then any linear operator L: X → Y is completely continuous (see also Theorem 0.16). For spaces whose dimension is infinite, the complete continuity differs from the continuity. Theorem 1.3. Let x1,x2, ... be linearly independent vectors in X a let Xn be the subspace of X generated by x1,...,xn. Then, there exists a sequence {yn}∞ n=1 such that yn = 1, yn ∈ Xn, n ∈ N, and inf x∈Xn−1 yn −x > 1 2 , n ≥ 2, n ∈ N. Proof. Since x1,x2, ... are linearly independent, xn /∈ Xn−1 and the distance between xn and Xn−1 is positive. Let us denote it by α and let x∗ be an element of Xn−1 for which xn −x∗ < 2α. Then, yn = xn −x∗ xn −x∗ , n ≥ 2, n ∈ N, because α = inf x∈Xn−1 xn −x = inf x∈Xn−1 xn −x∗ −x xn −x∗ . We add that we can easily put y1 = x1 x1 . Example 1.4. Let the dimension of X be infinite and let us consider the identical operator I on X. Using Theorem 1.3, in B[0,1], one can construct a sequence {yn}∞ n=1 such that yi −yn > 1 2 , i ∈ {1,2,...,n−1}, n ≥ 2, n ∈ N. Obviously, such a sequence cannot have a convergent subsequence. Therefore, B[0,1] is not (pre)compact and I is not completely continuous. Example 1.5. Let L be a continuous linear operator which maps X into a subspace of X with a finite dimension. The operator L is evidently completely continuous. Especially, in a Hilbert space, the (orthogonal) projection is completely continuous if and only if the considered subspace has a finite dimension. Note that an operator, which maps X into a subspace of X having a finite dimension, is called degenerated. 18 Example 1.6. In the space l2 , we consider the operator L: l2 → l2 given by x = {x1,x2,...,xn, ...} → Lx = x1, x2 2 ,..., xn 2n−1 , ... . This operator is completely continuous. It suffices to consider that the image of the unit ball is precompact and to use the linearity. Example 1.7. We consider C[a,b] with the norm f = max t∈[a,b] |f(t)|, f ∈ C[a,b], and the operator L: C[a,b] → C[a,b] defined by Lx = y(s) = b a k(s,t)x(t)dt, x ∈ C[a,b], s ∈ [a,b]. (1.1) It is possible to prove the following implication. If k is bounded for s ∈ [a,b], t ∈ [a,b] and all points of the discontinuity of k are on finitely many curves t = ϕk(s), k = {1,...,n}, where ϕk are continuous functions on [a,b], then the operator L given by (1.1) is completely continuous. We remark that this operator L is called the Fredholm operator. The requirement, that the points of the discontinuity of k are only on finitely many curves which intersect the lines s = const. in only one point, is essential. For example, for the function k(s,t) =    1, s < 1 2 ; 0, s ≥ 1 2 , the operator L maps x ≡ 1 into a discontinuous function. We prove the complete continuity of the operator L only in the case, when the function k is continuous on [a,b] × [a,b]. It is easy to see that Lx is defined correctly, Lx ∈ C[a,b], and that L is a linear and bounded operator (see also Example 0.5). We consider B[0,1] ⊆ C[a,b]. It suffices to show that the set L(B[0,1]) is precompact. We apply the Arzel`a–Ascoli theorem. Of course, L(B[0,1]) is a bounded set, because L is a bounded operator. It remains to show that L(B[0,1]) is a set of equicontinuous functions. For an arbitrarily given ε > 0, there exists δ > 0 such that |k(s1,t)−k(s2,t)| < ε if t ∈ [a,b], |s1 −s2| < δ, s1,s2 ∈ [a,b]. Then, |Lx(s1)−Lx(s2)| ≤ (b−a) max t∈[a,b] |k(s1,t)−k(s2,t)|· x ≤ ε(b−a) for all x ∈ B[0,1] and s1,s2 ∈ [a,b] satisfying |s1 −s2| < δ. If we put k(s,t) = 0 for t > s, then L takes the form Lx = y(s) = s a k(s,t)x(t)dt, x ∈ C[a,b], s ∈ [a,b]. (1.2) 19 If the function k is continuous, then the operator defined in (1.2) is completely continuous. This operator is called the Volterra operator. Remark 1.8. For a completely continuous operator, the image of the closed unit ball B[0,1] does not need to be compact, although it is precompact. As at the end of Example 1.7, on C[−1,1] with the uniform norm, we consider the completely continuous operator Jx(s) = s −1 x(t)dt, s ∈ [−1,1], x ∈ C[−1,1]. For n ∈ N, we put xn(t) =    0, −1 ≤ t ≤ 0; nt, 0 < t ≤ 1 n ; 1, 1 n < t ≤ 1. For n ∈ N, it is seen that xn ∈ C[−1,1], xn = 1, and that yn(s) = Jxn(s) =    0, −1 ≤ s ≤ 0; ns2 2 , 0 < s ≤ 1 n ; s− 1 2n , 1 n < s ≤ 1. We immediately see that, in C[−1,1], {yn}∞ n=1 converges to y(s) = 0, −1 ≤ s ≤ 0; s, 0 < s ≤ 1. But, for the operator J, the function y is not the image of any function from C[−1,1], because y is not continuous. However, it is possible to prove that, for any completely continuous linear operator, the image of B[0,1] is compact if the considered space is reflexive. 1.2 Basic properties Throughout this section, let X be a Banach space. Theorem 1.9. If {Ln}∞ n=1 is a sequence of linear completely continuous operators on X, which converges to an operator L: X → X, i.e., Ln −L → 0 as n → ∞, then L is completely continuous as well. Proof. It is sufficient to prove that, for an arbitrarily given bounded sequence {xk}∞ k=1 ⊂ X, one can extract a convergent subsequence from {Lxk}∞ k=1. Since the operator L1 is completely continuous, {L1xk}∞ k=1 has a convergent subsequence. Let x1 k ∞ k=1 be a subsequence of {xk}∞ k=1 such that L1x1 k ∞ k=1 converges. Now, let us consider L2x1 k ∞ k=1 . From this sequence, we can extract a convergent subsequence as well. Let x2 k ∞ k=1 be a subsequence of x1 k ∞ k=1 such that L2x2 k ∞ k=1 converges. We proceed in the same way. An the end, we consider the diagonal sequence xk k ∞ k=1 . Any of the operators L1,L2,L3, ... transforms this sequence into a 20 convergent one. We show that L also transforms it into a convergent sequence. Since X is complete, it suffices to show that Lxk k ∞ k=1 satisfies the Cauchy criterion. For n,k,l ∈ N, it holds Lxk k −Lxl l ≤ Lxk k −Lnxk k + Lnxk k −Lnxl l + Lnxl l −Lxl l . (1.3) Let ε > 0 be given and c > 0 be such that xk ≤ c, k ∈ N. Let n ∈ N be such that L−Ln < ε 3c . Then, we consider so large N ∈ N that Lnxk k −Lnxl l < ε 3 for all k,l > N, k,l ∈ N. Now, from (1.3), it follows that Lxk k −Lxl l < ε for all sufficiently large k,l. Remark 1.10. Since any linear combination of completely continuous operators is a completely continuous operator, from Theorem 1.9, we know that completely continuous linear operators form a closed subspace of L (X). Theorem 1.11. Let L1,L2 ∈ L (X) and let L1 be completely continuous. Then, the operators L1 ◦L2 and L2 ◦L1 are completely continuous as well. Proof. If a set M ⊆ X is bounded, then L2(M) = {y ∈ X; y = L2x,x ∈ M} is bounded as well. Thus, the set L1(L2(M)) is precompact and L1 ◦ L2 is completely continuous. If M ⊆ X is bounded, then L1(M) is precompact. Since L2 is continuous, L2(L1(M)) is precompact and L2 ◦L1 is completely continuous. Corollary 1.12. In the space X whose dimension is infinite, any linear completely continuous operator L: X → X does not have a bounded inverse L−1 . Proof. It is enough to consider Theorem 1.11 and the identical operator I = L ◦ L−1 (see Example 1.4). Theorem 1.13 (Schauder). The adjoint operator of a completely continuous operator L ∈ L (X) is completely continuous as well. Proof. We want to prove that L : X → X maps any ball into a precompact set. Due to the linearity of L , it suffices to show that the image L B of the closed unit ball with the centre in 0 ∈ X is precompact. We point out that B is the unit ball in X . Elements of X can be considered as functions defined on L(B[0,1]). We show that the set Φ of all functions assigned to the functionals belonging to the ball B is a set of uniformly bounded and equicontinuous functions. If a functional ϕ ∈ X satisfies ϕ ≤ 1, then sup x∈L(B[0,1]) |ϕ(x)| = sup x∈L(B[0,1]) |ϕ(x)| ≤ ϕ sup x∈B[0,1] Lx ≤ L 21 and ϕ(x )−ϕ(x ) ≤ ϕ · x −x ≤ x −x . Thus (consider the Arzel`a–Ascoli theorem), the set Φ is precompact in U = C L(B[0,1]) . But, the set Φ with the metric of the uniform convergence is isometric with the set L B with the metric given by the norm of X , because, for g1,g2 ∈ B , we have L g1 −L g2 = sup x∈B[0,1] L g1 −L g2,x = sup x∈B[0,1] | g1 −g2,Lx | = sup z∈L(B[0,1]) | g1 −g2,z | = sup z∈L(B[0,1]) | g1 −g2,z | = g1 −g2 . Since Φ is precompact, it is totally bounded. Hence, the set L B , which is isometric with it, is totally bounded as well. Therefore, the set L B is precompact in X . Remark 1.14. One can prove that the set Φ (from the proof of Theorem 1.13) is closed in U. Hence, Φ is compact and, consequently, the set L B is compact. By Remark 1.8, for a completely continuous linear operator, the image of the closed unit ball does not need to be compact. But, for any completely continuous linear operator on X , the image of the set B is compact. Theorem 1.15. Let L ∈ L (X) be a completely continuous operator. For arbitrary δ > 0, there exists only a finite number of linearly independent eigenvectors associated with eigenvalues of L whose absolute values are greater than δ. Proof. By contradiction, let us consider a sequence λ1,λ2,...λn, ... of eigenvalues of L such that |λn| > δ, n ∈ N, and a sequence x1,x2,...xn, ... of associated linearly independent eigenvectors. According to Theorem 1.3, we can construct a sequence y1,y2,...yn, ... such that yn = 1, yn ∈ Xn, n ∈ N, and that inf x∈Xn−1 yn −x > 1 2 , n ≥ 2, n ∈ N, where Xn is the subspace generated by x1,...,xn. The sequence {yn/λn}∞ n=1 is bounded, because |λn| > δ, n ∈ N. Now, we prove that, from the sequence of the images {L(yn/λn)}∞ n=1, one cannot choose a convergent subsequence. If yn = n ∑ k=1 αkxk, then L yn λn = n−1 ∑ k=1 αkλk λn xk +αnxn = yn +zn, where zn = n−1 ∑ k=1 αk λk λn −1 xk ∈ Xn−1. 22 Therefore, for any p,q ∈ N, q < p, it holds L yp λp −L yq λq = yp +zp −(yq +zq) = yp −(yq +zq −zp) > 1 2 , because yq +zq −zp ∈ Xp−1. We have a contradiction. Remark 1.16. Especially, from Theorem 1.15, it follows that the number of linearly independent eigenvectors associated with an eigenvalue λ = 0 of a completely continuous linear operator is finite. Remark 1.17. Let the dimension of X be infinite. For any completely continuous operator L ∈ L (X), we see that 0 ∈ σ(L). Indeed, if 0 /∈ σ(L), then L−1 is bounded on X, which is a contradiction with Theorem 1.11 (Corollary 1.12). In the case of a linear completely continuous operator, the spectrum is non-empty. Using the so-called Fredholm alternative, one can prove that σ(L) of a completely continuous operator L ∈ L (X) can contain only eigenvalues and 0. Thus, the spectrum of a completely continuous linear operator has a very simple structure. In particular, we recall that it is a closed set. Example 1.18. We consider the following series of examples. a) Let us consider the operator T : C[0,1] → C[0,1] given by T : f(t) → f t2 , t ∈ [0,1], f ∈ C[0,1]. See Example 0.41. Especially, we consider the norm from (0.1). The operator T is not completely continuous. The spectrum σ(T) = {λ ∈ C; |λ| = 1} is an uncountable set. In the case of a space whose dimension is infinite, we also know that 0 ∈ σ(L) for any completely continuous linear operator L. We can also use directly Definition 1.1 and consider the sequence {tn }∞ n=1 ⊆ C[0,1]. From the sequence {Ttn }∞ n=1 = t2n ∞ n=1 , one cannot choose a convergent subsequence. b) Let us consider the operator ˜T : C[0,1] → C[0,1] given by ˜T : f(t) → t 0 f(x)dx, t ∈ [0,1], f ∈ C[0,1]. See Example 0.42. Especially, we consider the norm from (0.1). We show that ˜T is completely continuous. We consider an arbitrary bounded sequence { fn}∞ n=1 ⊆ C[0,1] and, using the Arzel`a–Ascoli theorem, we prove that ˜T fn ∞ n=1 is precompact. Let K > 0 be such that fn ≤ K, n ∈ N. We have ˜T fn = max t∈[0,1] t 0 fn(x)dx ≤ max t∈[0,1] t 0 |fn(x)| dx ≤ 1 0 K dx = K, n ∈ N. 23 At the same time, for t,s ∈ [0,1], n ∈ N, we have ˜T fn(t)− ˜T fn(s) ≤ s t |fn(x)| dx ≤ K |t −s|. Thus, the sequence ˜T fn ∞ n=1 is uniformly bounded and equicontinuous. We remark that if we use the complete continuity of ˜T, then it is easy to determine the spectrum σ( ˜T) = {0} (cf. Example 0.42; see Remark 1.17). c) We consider L1,L2 ∈ L l2 from Example 0.18, f). For these operators, we determine σ(L1) and σ(L2) and we decide whether L1,L2 are completely continuous. We recall that L1x = 0,x1, x2 2 , x3 3 , ... , x = {xn}∞ n=1 ∈ l2 , L2x = {0,x1,x2,x3, ...}, x = {xn}∞ n=1 ∈ l2 . At first, we analyse L1. For a constant λ, we want to find x = {xn}∞ n=1 ∈ l2 such that L1x = λx. We get 0,x1, x2 2 , x3 3 , ... = {λx1,λx2,λx3, ...}. For λ = 0, we have x1 = 0, x2 = 0, x3 = 0, ...; and, for λ = 0, we also have x1 = x2 = x3 = ··· = 0. The equation L1x = λx does not have any non-zero solution for any λ. Therefore, any complex number is not an eigenvalue of the operator L1. We show that L1 is completely continuous. For all k ∈ N, we define Lk 1 : l2 → l2 by Lk 1 : x = {xn}∞ n=1 → 0,x1, x2 2 ,..., xk k ,0,0, ... . Any of the operators Lk 1 is linear, bounded, with a finite dimension of its range. Therefore, they are completely continuous. Since, for all k ∈ N and x = {xn}∞ n=1 ∈ l2 , x ≤ 1, it holds Lk 1x−L1x 2 = ∞ ∑ n=k+1 xn n 2 ≤ 1 k2 x 2 ≤ 1 k2 , we have Lk 1 −L1 2 ≤ 1 k2 . Hence, the completely continuous operators Lk 1 converge to L1 (as k → ∞), which proves that L1 is completely continuous (see Theorem 1.9). From Remark 1.17, we get that σ(L1) = {0}. Now, we prove that any λ satisfying |λ| ≤ 1 lies in σ(L2). In this case, the operator L2 − λI does not map l2 into l2 , because there is no z = {zn}∞ n=1 ∈ l2 such that (L2 −λI)z = {−λz1,z1 −λz2,z2 −λz3, ...} = {−1,0,0, ...}. 24 If we exclude the trivial case λ = 0, we have z = 1 λ , 1 λ2 , 1 λ3 , ... . Such z (for |λ| ≤ 1) is not in l2 . Therefore, the operator L2 is not completely continuous. It follows from the fact that its spectrum is the uncountable set {λ ∈ C; |λ| ≤ 1}. d) On the space C[0,1] with the uniform norm, we consider the operator given by F : f(t) → t2 f(0), t ∈ [0,1], f ∈ C[0,1]. This operator is completely continuous. Obviously, it is bounded, linear, and its range is a one-dimensional subspace of C[0,1]. The estimation |F f(t)| = t2 f(0) ≤ |f(0)| ≤ f , t ∈ [0,1], f ∈ C[0,1], gives F ≤ 1. Then, the choice f ≡ 1 shows that F = 1. The complete continuity of the operator F can be proved also directly. Let {fn}∞ n=1 ⊂ B[0,1] ⊂ C[0,1]. From the estimations F fn ≤ F · fn ≤ fn ≤ 1, n ∈ N, |F fn(t)−F fn(s)| = t2 −s2 fn(0) ≤ |(t −s)(t +s)| ≤ 2|t −s|, n ∈ N, t,s ∈ [0,1], and from the Arzel`a–Ascoli theorem, it follows the complete continuity of F. Now, we identify eigenvalues of the operator F. We look for non-trivial solutions of the equation t2 f(0) = λ f(t). For λ = 0, e.g., f(t) = t is a solution. Therefore, 0 is an eigenvalue. For λ = 0, we obtain f(t) = 1 λ f(0)t2 . Hence, f(0) = 0 and, consequently, f ≡ 0. Since F is completely continuous, σ(F) = {0}. e) On the complex space l2 , we define the operator R: l2 → l2 by Rx = {in xn}∞ n=1, x = {xn}∞ n=1 ∈ l2 , where i is the imaginary unit. For x = {xn}∞ n=1 ∈ l2 , we have Rx 2 = ∞ ∑ n=1 |xn|2 = x 2 , which gives R = 1. If λ is an eigenvalue, then there exists a non-zero element x = {xn}∞ n=1 ∈ l2 such that Rx = λx, i.e., in xn = λxn, n ∈ N. It is seen that the eigenvalues are i,i2 ,i3 ,i4 , i.e., i,−1,−i,1. For example, for i, the corresponding 25 eigenvector is {1,0,0, ...}. For λ /∈ {i,−1,−i,1}, the operator R − λI has an inverse. The inverse is Sx = 1 in −λ xn ∞ n=1 , x = {xn}∞ n=1 ∈ l2 , where Sx ≤ k x for k = max 1 |i−λ| , 1 |i2 −λ| , 1 |i3 −λ| , 1 |i4 −λ| . Note that the operator R is not completely continuous, because 0 /∈ σ(R) (the dimension of l2 is infinite). f) Let the operator T1 ∈ L L2 [0,1] be defined by T1 f(x) = x· 1 0 f(t)dt, x ∈ [0,1], f ∈ L2 [0,1]. For f ∈ L2 [0,1], we have T1 f = x· 1 0 f(t)dt =    1 0 x· 1 0 f(t)dt 2 dx    1 2 = 1 0 f(t)dt ·   1 0 x2 dx   1 2 = 1 √ 3 1 0 f(t)dt ≤ 1 √ 3   1 0 |f(t)|2 dt   1 2   1 0 1dt   1 2 = 1 √ 3 f . Therefore, T1 ≤ 1/ √ 3. For f ≡ 1, we see that f = 1, T1 f = 1/ √ 3. Thus, T1 = 1 √ 3 . We determine the discrete spectrum of T1. We want to find a non-trivial function f such that T1 f(x) = λ f(x), x ∈ [0,1]. For λ = 0, it suffices to consider an arbitrary identically non-zero function f ∈ C[0,1] for which 1 0 f(t)dt = 0. For example, f(x) = sin(2πx). For λ = 0 and T1 f(x) = x 1 0 f(t)dt = λ f(x), 26 we see that f has to be linear, i.e., f(x) = kx. Then, x 1 0 kt dt = λkx, x ∈ [0,1], i.e., λ = 1/2. Thus, the point spectrum is {0,1/2}. Obviously, the operator T1 is linear and continuous and its range has a finite dimension. Therefore, it is completely continuous and σ (T1) = {0,1/2}. Let the operator T2 ∈ L L2 [−1,1] be defined by T2 : f(x) → 1 −1 x2 t f(t)dt, x ∈ [−1,1], f ∈ L2 [−1,1]. For any f ∈ L2 [−1,1], we have T2 f 2 = 1 −1 |T2 f(x)|2 dx = 1 −1 x4   1 −1 t f(t)dt   2 dx = 2 5   1 −1 t f(t)dt   2 ≤ 2 5 t 2 · f 2 = 4 15 f 2 . For f(x) = x, we have f = 2 3 , T2 f 2 = 2 5 2 3 2 . Thus, T2 = 2 √ 15 . Let us find eigenvalues of T2. We look for non-trivial solutions of the equation (T2 −λI)f = 0. From the identity x2 1 −1 t f(t)dt = λ f(x), we see that the function f has to be a multiple of the function x2 for λ = 0, i.e., f(x) = kx2 . From x2 1 −1 tkt2 dt = λkx2 , it follows λ = 1 −1 t3 dt = 0. The point spectrum is {0}. Since T2 is completely continuous, σ(T2) = {0}. 27 Let the operator T3 ∈ L L2 [0,1] be defined by T3 f(x) = x· f(x), x ∈ [0,1], f ∈ L2 [0,1]. Since T3 f =   1 0 |x f(x)|2 dx   1 2 ≤   1 0 |f(x)|2 dx   1 2 = f , f ∈ L2 [0,1], we have the inequality T3 ≤ 1. Let us determine the discrete spectrum. We look for non-trivial solutions of the equation (T3 −λI) f = 0. As a solution of the equation (x − λ)f(x) = 0, x ∈ [0,1], we get only f = 0. Thus, the discrete spectrum is empty. Now, we consider the continuous spectrum. We consider a function g ∈ L2 [0,1] and find λ for which the equation (T3 − λI)f = g has a solution in L2 [0,1]. If λ = 0, then the equation x f(x) = g(x) has only the solution f(x) = g(x) x . For example, for g(x) = √ x ∈ L2 [0,1], we have f(x) = 1/ √ x /∈ L2 [0,1]. Thus, 0 ∈ σ(T3). In the case when λ = 0, we get the solution f(x) = g(x) x−λ , x ∈ [0,1]. For λ ∈ (0,1] and g ≡ 1, we obtain that f /∈ L2 [0,1]. For others λ ∈ C, i.e., λ /∈ [0,1], the function 1/(x−λ) is continuous on [0,1]. Therefore, the equation T3 f −λ f = g has a solution in L2 [0,1] for any function g ∈ L2 [0,1]. The spectrum is σ (T3) = [0,1]. The operator T3 cannot be completely continuous, because its spectrum is an uncountable set. Moreover, T3 = 1 (see Theorem 0.39). 1.3 Self-adjoint operator in Hilbert space For self-adjoint linear operators in Hilbert spaces with finite dimensions, we have the well-known theorem about the existence of an operator matrix in the diagonal form. Now, we extend this theorem to completely continuous self-adjoint operators in Hilbert spaces. Let H be a Hilbert space. Theorem 1.19. All eigenvalues λ of a self-adjoint operator L: H → H are real. Proof. Let Lx = λx for some non-zero x ∈ H. Then, λ x,x = λx,x = Lx,x = x,Lx = x,λx = ¯λ x,x . We see that λ = ¯λ. Theorem 1.20. Eigenvectors of a self-adjoint operator L: H → H, which correspond to different eigenvalues, are orthogonal. Proof. If Lx = λx and Ly = µy for λ = µ, then λ x,y = λx,y = Lx,y = x,Ly = x,µy = µ x,y , which gives x,y = 0. 28 Lemma 1.21. If a sequence {ξn}∞ n=1 ⊆ H and ξ ∈ H satisfy sup n∈N ξn < +∞ and L(ξn −ξ) → 0 as n → ∞ for a self-adjoint operator L: H → H, then Q(ξn) = Lξn,ξn → Lξ,ξ = Q(ξ) as n → ∞. Proof. For all n ∈ N, we have | Lξn,ξn − Lξ,ξ | ≤ | Lξn,ξn − Lξ,ξn |+| ξ,Lξn − ξ,Lξ | together with | Lξn,ξn − Lξ,ξn | = | L(ξn −ξ),ξn | ≤ ξn · L(ξn −ξ) and | ξ,Lξn − ξ,Lξ | = | ξ,L(ξn −ξ) | ≤ ξ · L(ξn −ξ) . Since the set of all numbers ξn for n ∈ N is bounded and L(ξn −ξ) → 0 as n → ∞, we get | Lξn,ξn − Lξ,ξ | → 0 as n → ∞. Lemma 1.22. If the functional ξ → |Q(ξ)| = | Lξ,ξ |, ξ ∈ H, where L: H → H is a self-adjoint operator, assumes a maximum on B[0,1] ⊆ H at an element ξ0, then ξ0,η = 0 implies that Lξ0,η = ξ0,Lη = 0. Proof. Obviously, ξ0 = 1. Let η = 0 and ξ0,η = 0. We set ξ = ξ0 +aη 1+|a|2 η 2 , where |a| > 0 is a sufficiently small number. From ξ0 = 1, we get ξ = 1. It holds Q(ξ) = 1 1+|a|2 η 2 Q(ξ0)+ ¯a Lξ0,η +a Lξ0,η +|a|2 Q(η) = 1 1+|a|2 η 2 Q(ξ0)+2 ¯a Lξ0,η +|a|2 Q(η) for such a number a that ¯a Lξ0,η is real. From the expression above, we see the following implication. If Lξ0,η = 0, then (consider |a| ≈ 0+ ) |Q(ξ)| > |Q(ξ0)|, which is a contradiction. 29 Remark 1.23. From Lemma 1.22, it follows that ξ0 is an eigenvector of the operator L if the functional |Q(ξ)| assumes a maximum on B[0,1] at ξ = ξ0. Theorem 1.24 (Hilbert–Schmidt). For every completely continuous self-adjoint operator L: H → H, there exists an orthonormal system of eigenvectors ϕ1,ϕ2, ... corresponding to non-zero eigenvalues λ1,λ2, ... such that any element ξ ∈ H can be uniquely written as ξ = N ∑ k=1 ckϕk + ¯ξ, where N ∈ N∪{∞}, ¯ξ satisfies L ¯ξ = 0, and Lξ = N ∑ k=1 λkckϕk. If the system of ϕk is infinite, then lim k→∞ λk = 0. Proof. By induction, we construct eigenvectors ϕn so that the absolute values of the corresponding eigenvalues satisfy |λ1| ≥ |λ2| ≥ ··· ≥ |λn| ≥ ··· In the construction of ϕ1, we investigate the functional |Q(ξ)| = | Lξ,ξ | and we prove that it assumes a maximum on B[0,1]. We denote S = sup ξ ≤1 | Lξ,ξ |. Let {ξk}∞ k=1 be a sequence such that ξk ≤ 1 for k ∈ N and | Lξk,ξk | → S as k → ∞. By Remark 1.14 (or Remark 1.8), from the sequence {ξk}∞ k=1, one can extract a subsequence {ξ1 k }∞ k=1 such that Lξ1 k −Lη → 0 as k → ∞ for some η ∈ H, where η ≤ 1. By Lemma 1.21, it holds | Lη,η | = S. We put ϕ1 = η. We add that η = 1. Indeed, for η < 1, the element η1 = η/ η satisfies η1 = 1, | Lη1,η1 | > S. We know that (see Remark 1.23) Lϕ1 = λ1ϕ1, where |λ1| = | Lϕ1,ϕ1 | ϕ1,ϕ1 = | Lϕ1,ϕ1 | = S. Let eigenvectors ϕ1,ϕ2,...,ϕn correspond to eigenvalues λ1,λ2,...,λn from our construction. Let Hn be the subspace of H generated by ϕ1,ϕ2,...,ϕn. We consider the functional | Lξ,ξ | on the set H⊥ n ∩ B[0,1] ⊆ H. Since the subspace Hn is invariant and L is self-adjoint, the set H⊥ n is invariant with respect to L (see Remark 0.36). According to the considerations above for H⊥ n , we get that, in H⊥ n ∩ B[0,1], one can find the required element (denoted as ϕn+1) which is an eigenvector of the operator L. The following two cases are possible: 30 1. after a finite number of the steps, we get a subspace H⊥ n0 , where Lξ,ξ ≡ 0; 2. Lξ,ξ ≡ 0 on H⊥ n for all n ∈ N. In the first case, from Lemma 1.22, it follows that L maps the subspace H⊥ n0 into {0}, i.e., H⊥ n0 is composed of eigenvectors corresponding to the eigenvalue λ = 0. In this case, the set {ϕk} is finite. In the second case, we obtain a sequence {ϕk}∞ k=1 of eigenvectors for which λk = 0, k ∈ N. We know that λk → 0 as k → ∞ (see Theorem 1.15). Let ¯H = ∞ n=1 H⊥ n = {0}. If ξ ∈ ¯H, then | Lξ,ξ | ≤ |λn|· ξ 2 , n ∈ N, i.e., Lξ,ξ = 0. Therefore, by Lemma 1.22 (for the subspace ¯H), the operator L maps ¯H into {0}. From the construction of {ϕk}N k=1, it follows that any element ξ ∈ H can be uniquely expressed as ξ = N ∑ k=1 ckϕk + ¯ξ, where N ∈ N∪{∞}, L ¯ξ = 0, and Lξ = N ∑ k=1 λkckϕk. Remark 1.25. Theorem 1.24 says that, for any completely continuous self-adjoint operator on H, there exists an orthogonal base of the space H which is composed of eigenvectors of this operator. To obtain such a base, it is enough to consider {ϕk}N k=1 with an arbitrary orthogonal base of H⊥ n0 or ¯H (see the proof of Theorem 1.24). In other words, we get a result entirely analogous to the theorem about the existence of an operator matrix in the diagonal form for self-adjoint operators in a space with a finite dimension. 31 2. Derivative in Banach spaces Let X,Y be real Banach spaces. 2.1 Weak and strong derivative Let f be a map defined on an open set G ⊆ X with values in Y. Definition 2.1. For x ∈ G, we define the directional derivative of f in the direction of h ∈ X as the limit lim λ→0 f(x+λh)− f(x) λ if it exists. We denote it by Dh f(x). Definition 2.2. If f has the directional derivative in the direction of any h ∈ X at x ∈ G and df(x): h → Dh f(x) is a continuous linear map from X to Y, we say that f has the weak derivative d f(x) at x. The weak derivative is sometimes called the Gâteaux derivative. Definition 2.3. If there exists a continuous linear map L: X → Y such that lim h→0 f(x+h)− f(x)−L(h) h = 0, (2.1) we say that f has the strong derivative at x ∈ G. If f has the strong derivative at x ∈ G, the map L from (2.1), which is uniquely determined, is called the Fréchet derivative of f at x and is denoted by f (x). Remark 2.4. If f (x) exists, f has also the weak derivative at x and f (x) = d f(x). In this case, there exists the directional derivative of f in the direction of any h and Dh f(x) = f (x)(h). Remark 2.5. Let f be a real function on X (i.e., G = X and f is a functional). The function f has the Gâteaux derivative at x ∈ X if there exists L ∈ X such that lim λ→0 f(x+λh)− f(x) λ = Lh (2.2) for all h ∈ X. Example 2.6. On the Banach space X = C[0,1] (with the norm of the uniform convergence), we consider the functional F : f → 1 0 f2 . 32 For any ϕ ∈ X, we compute the directional derivative of F in the direction of ϕ as (see (2.2)) Dϕ F(f) = lim λ→0 F(f +λϕ)−F(f) λ = lim λ→0 1 λ 1 0 (f +λϕ)2 − f2 = lim λ→0 1 0 2 fϕ +λϕ2 = 1 0 2 fϕ. We introduce L: ϕ → 1 0 2 fϕ. It is immediately seen that L is a linear functional on X. From |L(ϕ)| ≤ 2 1 0 |fϕ| ≤ 2 max t∈[0,1] |ϕ(t)|· 1 0 |f| = 2 ϕ 1 0 |f|, it follows that L is bounded. Thus, F has the Gâteaux derivative at f (which is given by L). If the Fréchet derivative of F at f exists, then it is L (see Remark 2.4). Its existence is guaranteed by the limit lim ϕ→0 F(f +ϕ)−F(f)−L(ϕ) ϕ = lim ϕ→0 1 ϕ 1 0 (f +ϕ)2 − f2 −2 fϕ = lim ϕ→0 1 ϕ 1 0 ϕ2 ≤ lim ϕ→0 ϕ = 0. We see that the functional F has also the strong derivative at f and its Fréchet derivative is equal to L. Before the next example, we recall that continuous linear functionals on l1 have the form ∞ ∑ n=1 anxn for {an}∞ n=1 ∈ l∞ . If {an}∞ n=1 ∈ l∞ and Ψ(x) = ∞ ∑ n=1 anxn for x = {xn}∞ n=1 ∈ l1 , then Ψ ∈ l1 (= l∞ ). Example 2.7. In the Banach space l1 , we consider the function f : t → t on l1 and x = {xn}∞ n=1 ∈ l1 . We show that f has the Gâteaux derivative at x if and only if xn = 0 for all n ∈ N. In this case, df(x) = {sgnxn}∞ n=1 ∈ l∞ . 33 Then, we show that the Fréchet derivative of the norm in l1 does not exist at any point. Let {xn}∞ n=1 be a sequence in l1 such that xk = 0 for some k ∈ N. We put ek = {0,...,0,1,0, ...}, where 1 is in the k-th position. The non-existence of the directional derivative of f in the direction of ek comes from x+λek − x λ = 1 λ ∞ ∑ n=1 |xn|+|λ|− ∞ ∑ n=1 |xn| = |λ| λ . For λ → 0, we see that the limit of the considered term does not exist. Therefore, the function f does not have the weak (and also the strong) derivative at any point with a zero element. Let {xn}∞ n=1 be a sequence in l1 such that xn = 0 for all n ∈ N. We consider h = {hn}∞ n=1 ∈ l1 and ε > 0. Let k ∈ N be such that ∞ ∑ n=k+1 |hn| < ε. Obviously, there exists δ > 0 such that sgn(xn +λhn) = sgnxn, |λ| < δ, n ∈ {1,...,k}. For λ ∈ (−δ,δ), we obtain x+λh − x λ − ∞ ∑ n=1 hn sgnxn = 1 λ k ∑ n=1 |xn +λhn|−|xn|−λhn sgnxn + ∞ ∑ n=k+1 |xn +λhn|−|xn|−λhn sgnxn ≤ 1 |λ| ∞ ∑ n=k+1 (|xn|+|λ|·|hn|−|xn|+|λ|·|hn|) ≤ 1 |λ| ∞ ∑ n=k+1 2|λ|·|hn| < 2ε. Due to the fact that ε > 0 is arbitrary, the function f has the weak derivative at x which is equal to {sgnxn}∞ n=1 ∈ l∞ , because the map h → ∞ ∑ n=1 hn sgnxn is a continuous linear functional on l1 . Now, we show that f does not have the strong derivative at any point. We assume that f has the derivative f (x) at x = {xn}∞ n=1. If f (x) exists, it is {sgnxn}∞ n=1 ∈ l∞ . We consider hj = 0,0,...,0,−2xj,−2xj+1,−2xj+2, ... , j ∈ N, where the value −2xj is in the j-th position. Obviously, hj = 2 ∞ ∑ n=j |xn| → 0 as j → ∞ 34 and (see (2.1)) x+hj − x − f (x) hj = x+hj − x − ∞ ∑ n=j (−2xn)sgnxn = ∞ ∑ n=1 |xn|− ∞ ∑ n=1 |xn|+ ∞ ∑ n=j 2|xn| = hj . Now, it is seen that f cannot have the strong derivative at x. 2.2 Convex function Now, we study derivatives of convex functions. Definition 2.8. A real function f : D ⊆ X → R is called convex on a convex set D ⊆ X if f(λx+(1−λ)y) ≤ λ f(x)+(1−λ)f(y), x,y ∈ D, λ ∈ [0,1]. An elementary example of convex functions on a Banach space is the norm. Theorem 2.9. Let f be a real convex function on an open convex set D ⊆ X which is continuous at x0 ∈ D. Then, there exist K > 0 and δ > 0 such that |f(x)− f(y)| ≤ K x−y , x,y ∈ B(x0,δ), where B(x0,δ) is the open ball with the center in x0 and the radius δ. Proof. The continuity of f at x0 guarantees the existence of M > 0 and δ > 0 such that |f(t)| ≤ M for t ∈ B(x0,2δ) ⊆ D. We consider x,y ∈ B(x0,δ), where x = y. We denote α = x−y , z = y+ δ α (y−x). We see that z ∈ B(x0,2δ). Since y is a convex combination of z and x, where y = α α +δ z+ δ α +δ x, using the convexity of f, we have f(y)− f(x) ≤ α α +δ f(z)+ δ α +δ f(x)− f(x) = α α +δ (f(z)− f(x)) ≤ α α +δ 2M ≤ 2M δ x−y . Analogously, f(x)− f(y) ≤ 2M δ y−x . Remark 2.10. On spaces whose dimension is infinite, there exist discontinuous linear functionals and all linear functional is a convex function. Thus, there exist discontinuous convex functions. Note that convex functions on open subsets of a space with a finite dimension are continuous. 35 Remark 2.11. From the proof of Theorem 2.9, it follows that it suffices to assume only the boundedness of f on some neighbourhood of x0. Thus, we know that a convex function is continuous on an open convex set D ⊆ X if and only if it is locally bounded on D. Before the following theorem, we recall that a real function f defined on a metric space M is called upper semi-continuous on M if the set {x ∈ M; f(x) < α} is open for all α ∈ R. Theorem 2.12. Let f be a convex function on an open convex set D ⊆ X. Then, the following conditions are equivalent: i) f is continuous on D; ii) f is upper semi-continuous on D; iii) f is upper bounded on some neighbourhood of a point of D; iv) f is continuous at a point of D. Proof. The implication i) ⇒ ii) is obvious. If a real function f is upper semi-continuous on D and x ∈ D, then the set {t ∈ D; f(t) < f(x)+1} is a neighbourhood of x, where f is upper bounded. We have proved the implication ii) ⇒ iii). For the implication iii) ⇒ iv), we consider that f is upper bounded on a neighbourhood of x0 ∈ D. Let f ≤ K on B(x0,δ) and let t ∈ B(x0,δ). We have 2x0 −t ∈ B(x0,δ). Thus, f(x0) = f 2x0 −t 2 + t 2 ≤ 1 2 f(2x0 −t)+ 1 2 f(t) ≤ 1 2 (K + f(t)). Then, we have −f(t) ≤ K −2 f(x0) ≤ K +2|f(x0)|. Since also f(t) ≤ K ≤ K +2|f(x0)|, we obtain |f(t)| ≤ K +2|f(x0)|, t ∈ B(x0,δ). Therefore, f is bounded on B(x0,δ) and it suffices to use Remark 2.11. Now, we consider the implication iv) ⇒ i). Let f be continuous at x ∈ D and let y ∈ D (y = x) be arbitrarily given. There exist z ∈ D and λ ∈ (0,1) such that y = λx + (1 − λ)z. We consider δ > 0 and K > 0 such that f ≤ K on B(x,δ) ⊆ D. We show that f is upper bounded on the set B(y,λδ) which implies that the function f is continuous at y. We consider t ∈ B(y,λδ). Since t = λ x+ t −y λ +(1−λ)z ∈ D, 36 where x+ t −y λ ∈ B(x,δ), the convexity of f gives the estimation f(t) ≤ λ f x+ t −y λ +(1−λ)f(z) ≤ λK +(1−λ)f(z). Remark 2.13. It is known that linear functionals are continuous if and only if they are continuous at 0, which is if and only if they are bounded. Continuous linear functionals are bounded on the unit ball, but continuous convex functions do not need to be bounded on the unit ball (although they are locally bounded). In addition, on any separable Banach space whose dimension is infinite, there exists continuous convex function, which is not bounded on the unit ball. Before the following theorem, we recall that a real function p on X is called a convex functional on X if: a) p(λx) = λ p(x), λ ≥ 0, x ∈ X; b) p(x+y) ≤ p(x)+ p(y), x,y ∈ X. Theorem 2.14. Let D ⊆ X be an open convex set, let f be a convex function on D, and let x ∈ D. Then, d+ f(x)(h) = lim t→0+ f(x+th)− f(x) t exists for all h ∈ X and the map d+ f(x): h → d+ f(x)(h) is a convex functional on X. Proof. Let h ∈ X. The function t → f(x+th)− f(x) t is non-decreasing on a right neighbourhood of 0. If t,s, where 0 < t < s, are sufficiently small (so that x+sh ∈ D), then we have x+th = s−t s x+ t s (x+sh) and, consequently, f(x+th) ≤ s−t s f(x)+ t s f(x+sh). Thus, 1 t (f(x+th)− f(x)) ≤ 1 s (f(x+sh)− f(x)). If we consider t > 0 sufficiently small, then − f(x−2th)− f(x) 2t ≤ f(x+2th)− f(x) 2t , (2.3) because 2 f(x) = 2 f x−2th+x+2th 2 ≤ f(x−2th)+ f(x+2th). (2.4) 37 Therefore (see (2.3)), we have −d+ f(x)(−h) ≤ d+ f(x)(h). Especially, the considered limit exists. It remains to prove the convexity of the functional. If λ > 0, then d+ f(x)(λh) = λ lim t→0+ f(x+tλh)− f(x) λt = λd+ f(x)(h). For h,k ∈ X, it holds (see (2.4)) d+ f(x)(h+k) = lim t→0+ f(x+(h+k)t)− f(x) t ≤ lim t→0+ f(x+2th)− f(x) 2t + f(x+2tk)− f(x) 2t = d+ f(x)(h)+d+ f(x)(k). Remark 2.15. In the proof of the previous theorem, we have obtained the inequality −d+ f(x)(−h) ≤ d+ f(x)(h). It is easy to show that Dh f(x) exists for all h ∈ X if and only if −d+ f(x)(−h) = d+ f(x)(h), h ∈ X. Theorem 2.16. Let f be a convex function on an open convex set D ⊆ X, let it be continuous at x ∈ D, and let it have the derivative Dh f(x) linearly in the direction of all h ∈ X. Then, f has the Gâteaux derivative d f(x) at x. Proof. By Theorem 2.9, there exist K > 0 and δ > 0 such that |f(u)− f(v)| ≤ K u−v for all u,v ∈ B(x,δ) ⊆ D. We consider h ∈ X. Let λ > 0 be such that x+λh ∈ B(x,δ). Then, |f(x+λh)− f(x)| ≤ K λh = Kλ h . Therefore, |d+ f(x)(h)| ≤ K h and, consequently, the derivative d+ f(x)(h) = Dh f(x) is continuous. 2.3 Tangent functional We know that the function f : X → R given by x → x is continuous and convex. We show that, under certain assumptions, the weak derivative of this function f is the so-called tangent functional. We recall the well-known corollary of the Hahn–Banach theorem which says that, for any non-zero element x ∈ X, there exists g ∈ X such that g = 1 and g(x) = x . This functional is called the tangent functional at x. 38 Theorem 2.17. Let the function f : x → x have the Gâteaux derivative at x = 0, x ∈ X. Then, the Gâteaux derivative d f(x) ∈ X has the norm d f(x) = 1 and d f(x)(x) = x . At the same time, if g ∈ X satisfies g = 1 and g(x) = x , then g = d f(x). Proof. For h ∈ X and sufficiently small t = 0, from the estimation 1 t [ x+th − x ] ≤ 1 |t| x+th−x = h , it follows that d f(x) ≤ 1. Since d f(x) x x = lim t→0 1 t x+t x x − x = lim t→0 1 t 1+ t x x − x = lim t→0 x t 1+ t x −1 = 1, we have d f(x) = 1 and d f(x)(x) = x . Let g ∈ X be such that g = 1 and g(x) = x . We consider h ∈ X and we define ε(t) = df(x)(h)− 1 t [ x+th − x ] for sufficiently small t = 0. Obviously, ε(t) → 0 as t → 0. For the considered t, we have g(x+th) ≤ g · x+th = x +td f(x)(h)−tε(t) = g(x)+td f(x)(h)−tε(t). Therefore, tg(h) ≤ td f(x)(h)−tε(t) which gives (as t → 0+ ) g(h) ≤ d f(x)(h). Since this inequality is valid for all h ∈ X (also for −h), it is enough to consider the linearity of g and df(x). Therefore, g = d f(x). The previous theorem relates to the geometry of a Banach space. According to this theorem, at all non-zero point in which the norm has the Gâteaux derivative, the space is “smooth” in the sense that there exists just one tangent functional. We discuss this topic in the next chapter. 39 3. Strictly and uniformly convex spaces In this chapter, we consider a real Banach space X. 3.1 Strictly convex space Definition 3.1. An extreme point of a convex set C ⊆ X is a point x ∈ C for which x = (a+b)/2, where a,b ∈ C, implies a = b. The set of all extreme points of the set C is denoted by ext C. Remark 3.2. Definition 3.1 says that x ∈ C is an extreme point if there does not exist a non-degenerated line segment in C having the center x. It is seen that x is an extreme point of C if and only if the set C {x} is convex. Definition 3.3. A Banach space is called strictly convex if all point of the unit sphere ∂B(0,1) is an extreme point of the closed unit ball B[0,1], i.e., if ext B[0,1] = ∂B(0,1). Strictly convex spaces are also called rotund. Remark 3.4. In strictly convex spaces, any line segment cannot lie on any sphere. Therefore, the following implication is valid. If C is a convex set in the strictly convex space X, x ∈ X, a,b ∈ C, and if x−a = x−b = dist(x,C), then a = b. Indeed, for λ ∈ [0,1], it suffices to consider that x−λa−(1−λ)b ≤ λx−λa + (1−λ)x−(1−λ)b = dist(x,C). Remark 3.5. Since extreme points of the closed unit ball B[0,1] have to lie on the unit sphere ∂B(0,1), we can say that X is strictly convex if and only if x+y < 2, x,y ∈ B[0,1], x = y. Let the identity x+y 2 = 2 x 2 +2 y 2 (3.1) be valid for points x,y of a Hilbert space. Since, in any Hilbert space, the identity x+y 2 + x−y 2 = 2 x 2 +2 y 2 is valid, we see that x = y. Geometrically, (3.1) says that, in the parallelogram with the sides x and y, the second diagonal x −y is missing. Now, we consider points x,y of X for which (3.1) is valid. Since 0 = 2 x 2 +2 y 2 − x+y 2 ≥ 2 x 2 +2 y 2 −( x + y )2 = 2 x 2 +2 y 2 − x 2 −2 x · y − y 2 = ( x − y )2 ≥ 0, we get x = y , but not x = y necessarily. This observation motivates the following theorem. 40 Theorem 3.6. For the Banach space X, the following statements are equivalent: i) X is strictly convex; ii) if x+y = x + y , x = 0,y = 0, then x = λy for some λ > 0; iii) if x,y ∈ X and x+y 2 = 2 x 2 +2 y 2 , then x = y, Proof. We begin with the implication i) ⇒ ii). Let X be strictly convex and let x,y be non-zero points of X satisfying the equality x+y = x + y . For example, let x ≤ y . We have x x + y y ≥ x x + y x − y x − y y = 1 x ( x + y )− y 1 x − 1 y = 2. Therefore, we obtain (see Remark 3.5) x x = y y , i.e., x = x y y. Let us consider the implication ii) ⇒ iii). Let the implication in ii) be true and let x+y 2 = 2 x 2 +2 y 2 . Hence (see the text before Theorem 3.6), x = y . We do not consider the trivial case x = 0,y = 0. In the non-zero case, we obtain x+y 2 = 4 x 2 . Therefore, x+y = 2 x = x + y . Based on the assumption, x = λy for some λ > 0. Considering x = y , we obtain that λ = 1. The implication has been proved. In the last part of the proof, we can use, e.g., Remark 3.4 or Remark 3.5. If x = y = 1 2 (x+y) = 1, then x+y 2 = 4 = 2 x 2 +2 y 2 . 41 3.2 Uniformly convex space If points x,y of a strictly convex space have a constant distance and they are on the unit sphere ∂B(0,1), then the midpoint of the line segment between x and y is in the open unit ball. But, in a space whose dimension is infinite, it is not clear, whether this midpoint can be arbitrarily close to the sphere ∂B(0,1). This motivates the following definition. Definition 3.7. A Banach space is called uniformly convex if, for all ε ∈ (0,2], there exists δ > 0 such that, if x,y are in the unit ball B[0,1] and x−y ≥ ε, then 1 2 (x+y) ≤ 1−δ. The basic characteristics of uniformly convex spaces are mentioned in the following theorem. Theorem 3.8. For the Banach space X, the following statements are equivalent: i) X is uniformly convex; ii) for any ε ∈ (0,2], there exists δ > 0 such that, if x,y are on the unit sphere ∂B(0,1) and x−y ≥ ε, then 1 2 (x+y) ≤ 1−δ; iii) if {xn}∞ n=1 ,{yn}∞ n=1 ⊆ ∂B(0,1) satisfy lim n→∞ xn +yn 2 = 1, then xn −yn → 0 as n → ∞. Proof. If X is uniformly convex, then ii) is true. The implication ii) ⇒ iii) is trivial as well. If X is not uniformly convex, then there exist ε > 0 and sequences {xn}∞ n=1 ,{yn}∞ n=1 ⊆ B[0,1] such that xn −yn ≥ ε, n ∈ N, and 1− 1 n ≤ 1 2 xn +yn ≤ 1 2 ( xn + yn ) ≤ 1, n ∈ N. Therefore, xn + yn → 2 as n → ∞. Since xn ≤ 1, yn ≤ 1, n ∈ N, we have lim n→∞ xn = lim n→∞ yn = 1. Especially, without loss of generality, we can assume that xn · yn > 0 for all n ∈ N. For un = xn xn , vn = yn yn , n ∈ N, 42 we obtain un = 1 = vn , n ∈ N, and | un +vn − xn +yn | ≤ un −xn + vn −yn , n ∈ N. The right side converges to 0 as n → ∞, because un −xn = xn xn −xn = xn · 1 xn −1 = 1− xn , n ∈ N, and, analogously, vn −yn = 1− yn , n ∈ N. With respect to xn +yn 2 → 1 as n → ∞, we get un +vn 2 → 1 as n → ∞. (3.2) We have two sequences {un}∞ n=1 ,{vn}∞ n=1 ⊆ ∂B(0,1) for which (3.2) is valid and, at the same time, for which liminf n→∞ un −vn ≥ ε > 0, because 0 < ε ≤ xn −yn ≤ xn −un →0 + un −vn + vn −yn →0 , n ∈ N. Remark 3.9. Many other equivalences can be mentioned in Theorems 3.6 and 3.8. The statements in Theorem 3.6 are also equivalent to: iv) if p ∈ (1,∞) and x,y ∈ X, x = y, then x+y 2 p < x p + y p ; v) if x−y = x−z + z−y , then there exists λ ∈ [0,1] such that z = λx+(1−λ)y. Similarly, the statements in Theorem 3.8 are equivalent to: iv) for any ε > 0, there exists δ > 0 such that, if x < 1+δ, y < 1+δ, and if 1 2 (x+y) ≥ 1, then x−y < ε. Remark 3.10. Obviously, any uniformly convex space is strictly convex. In spaces whose dimension is finite, these notions are same. It follows from the compactness of the closed unit ball B[0,1] in spaces with finite dimensions and the continuity of the function (x,y) → x+y 2 . 43 Remark 3.11. We consider X = C[0,1] with the norm f = max{|f(t)|; t ∈ [0,1]}+   1 0 f2   1 2 , f ∈ X. One can show that this space X is strictly convex, but it is not uniformly convex. Example 3.12. Any Hilbert space H is uniformly convex. It is enough to consider that, for x,y ∈ H, x ≤ 1, y ≤ 1, and x−y ≥ ε, we have x+y 2 2 + x−y 2 2 = x 2 2 + y 2 2 and, consequently, x+y 2 2 ≤ 1 2 + 1 2 − ε 2 2 , i.e., x+y 2 ≤ 1− ε2 4 < 1. Example 3.13. The spaces Zp = Lp (Ω), · p are strictly convex for all p ∈ (1,∞) and any measurable set Ω ⊆ R. This fact is possible to easily show taking into account the case, when the Minkowski inequality becomes the equality. But, we prove the stronger result that the space Zp is uniformly convex for all p > 1. It is enough to prove that, for all ε > 0, there exists δ > 0 such that, if u,v ∈ Lp (Ω), u p = v p = 1, and if u+v 2 p p > 1−δ, then u−v 2 p p ≤ 2εp . We consider an arbitrary number ε > 0. For simplicity, we denote s = 1 2 (u+v), t = 1 2 (u−v), where u = s+t, v = s−t. We put S = {ω ∈ Ω; |t(ω)| ≤ ε|s(ω)|}, S0 = {ω ∈ Ω; t(ω) = 0}, S+ = {ω ∈ Ω; 0 < |t(ω)| ≤ ε|s(ω)|}, T = {ω ∈ Ω; |t(ω)| > ε|s(ω)|}. Evidently, S |t(ω)|p dω ≤ εp Ω |s(ω)|p dω ≤ εp . (3.3) 44 It is well-known that the function λ → |λ|p is strictly convex and continuous on R. Therefore, |λ +1|p +|λ −1|p 2 > |λ|p = λ +1 2 + λ −1 2 p , λ ∈ R, and there exists γ > 0 such that 1 2 [|λ +1|p +|λ −1|p ]−|λ|p ≥ γ, λ ∈ − 1 ε , 1 ε . (3.4) In (3.4), we consider λ = s(ω) t(ω) , ω ∈ Ω S0. For ω ∈ T, we obtain 1 2 [|s(ω)+t(ω)|p +|s(ω)−t(ω)|p ] ≥ γ|t(ω)|p +|s(ω)|p and, for ω ∈ S = S0 ∪S+, it holds 1 2 [|s(ω)+t(ω)|p +|s(ω)−t(ω)|p ] ≥ |s(ω)|p . Thus, 1 = Ω 1 2 [|s(ω)+t(ω)|p +|s(ω)−t(ω)|p ] dω ≥ Ω |s(ω)|p dω + T γ|t(ω)|p dω. (3.5) If Ω |s(ω)|p dω > 1−δ, then, using (3.5), we obtain T |t(ω)|p dω ≤ δ γ . (3.6) The choice δ = γεp , (3.3), and (3.6) give Ω |t(ω)|p dω = T |t(ω)|p dω + S |t(ω)|p dω ≤ εp +εp = 2εp . 3.3 Projection Now, we focus on projections in uniformly convex Banach spaces. Let X be a uniformly convex Banach space. Theorem 3.14. Let C(= /0) be a closed convex subset of X. Then, for any y ∈ X, there exists just one c ∈ C such that y−c = dist(y,C). 45 Proof. The uniqueness is presented in Remark 3.4. Hence, it suffices to prove the existence of an element x ∈ C with the minimal norm, because, without loss of generality, we can assume that y = 0. We put d = inf{ c ; c ∈ C}. If d = 0, then we see that 0 ∈ C (the set C is closed). Therefore, without loss of generality, let d = 1. In this case, there exists a sequence {xn}∞ n=1 ⊆ C such that lim n→∞ xn = 1. If {xn}∞ n=1 is Cauchy, then there exists the limit x = lim n→∞ xn, where x ∈ C, x = 1. The convexity of C guarantees that 1 2 (xn +xk) ≥ 1, n,k ∈ N. We consider an arbitrary number ε > 0. For the given ε, there exists n0 ∈ N such that xn < 1 +ε for all n ≥ n0, n ∈ N. Therefore (using the uniform convexity), we have xn −xk < ξ(ε) for all n,k ≥ n0, n,k ∈ N, where ξ(ε) → 0 as ε → 0+ . The theorem is proved. Definition 3.15. Let C be a closed convex subset of X. For any x ∈ X, the uniquely determined PC(x) ∈ C satisfying x−PC(x) = dist(x,C) is called the projection (of x on C). Theorem 3.16. Let C be a closed convex subset of X. The projection PC is continuous. Proof. For simplicity, we consider dist(0,C) = 1. Let {xn}∞ n=1 ⊆ X be a sequence such that xn → 0 as n → ∞. We want to show that PC(xn) → PC(0) as n → ∞. Since | xn −PC(xn) −1| = |dist(xn,C)−dist(0,C)| ≤ xn −0 = xn → 0 as n → ∞, we obtain that xn −PC(xn) → 1 as n → ∞. From 1 ← | xn −PC(xn) − xn | ≤ PC(xn) ≤ xn −PC(xn) + xn → 1 as n → ∞, we have PC(xn) → 1 as n → ∞. The convexity of C gives 1 ≤ 1 2 (PC(0)+PC(xn)) ≤ 1 2 PC(0) + 1 2 PC(xn) → 1 as n → ∞. We use iii) from Theorem 3.8 for the sequences {PC(0)}∞ n=1 , PC(xn) PC(xn) ∞ n=1 . Thus, PC(xn)−PC(0) → 0 as n → ∞. It suffices to consider that 1 2 (PC(0)+PC(xn)) − 1 2 PC(0)+ PC(xn) PC(xn) ≤ 1 2 PC(xn)− PC(xn) PC(xn) → 0 as n → ∞. 46 Remark 3.17. The projection on closed convex sets does not need to be linear (even in Hilbert spaces). It is known that the projection on (closed) subspaces of Hilbert spaces is linear. But, it is not true in uniformly convex spaces. For a better understanding of Theorem 3.20 below, we mention the following result. Theorem 3.18. Any uniformly convex space is reflexive. Remark 3.19. Strictly convex spaces do not need to be reflexive (see the example in Remark 3.11). Now, without a proof, we mention a generalization of Theorem 3.14. Theorem 3.20. Let C be a closed convex subset of a strictly convex reflexive Banach space Y and y ∈ Y. Then, there exists just one c ∈ C such that y−c = dist(y,C). Remark 3.21. We add that there exist strictly convex reflexive spaces, which are not uniformly convex. Now, we generalize the concept of the projection to the concept of the so-called metric projections in the following definition. Definition 3.22. Let M be a subset of a Banach space Y. For x ∈ Y, we denote PM(x) = {m ∈ M; x−m = dist(x,M)}. The set M is called: • proximinal if PM(x) = /0 for all x ∈ Y; • semi-Chebyshev if PM(x) has at most one element for all x ∈ Y; • Chebyshev if PM(x) has just one element for all x ∈ Y. We repeat that, in strictly convex spaces, any closed convex set is semi-Chebyshev and that any closed convex subset of a strictly convex reflexive Banach space is Chebyshev (see Remark 3.4 and Theorem 3.20). We add that any compact set is proximinal. 3.4 Smooth space Now, let us define smooth spaces explicitly. Definition 3.23. The space X is called smooth at x ∈ ∂B(0,1) if there exists just one functional ϕ ∈ X such that ϕ = 1, ϕ(x) = 1. We say that X is smooth if it is smooth at any point of the unit sphere ∂B(0,1). We emphasize that, according to the above mentioned corollary of the Hahn–Banach theorem, the tangent functional exists at any point x ∈ ∂B(0,1). Concerning Definition 3.23, we point out the uniqueness. The following reinforcement of Theorem 2.17 is valid. Theorem 3.24 (Šmuljan). The space X is smooth at x ∈ ∂B(0,1) if and only if the function f : t → t has the Gâteaux derivative at x. The most important connection between Chapters 2 and 3 is presented in the following theorem. 47 Theorem 3.25 (Klee). If X is strictly convex, then X is smooth. If X is smooth, then X is strictly convex. Proof. We assume that X is not smooth at x ∈ ∂B(0,1). Thus, there exist ϕ,ψ ∈ X such that ϕ = ψ, ϕ = ψ = 1 = ϕ(x) = ψ(x). Since ϕ +ψ 2 = 1, the space X cannot be strictly convex. If X is not strictly convex, then there exist x,y ∈ ∂B(0,1) such that x = y and x+y 2 ∈ ∂B(0,1). By the recalled corollary of the Hahn–Banach theorem, there exists a functional ϕ such that ϕ = 1, ϕ x+y 2 = 1. Since 1 = ϕ x+y 2 = 1 2 ϕ(x)+ 1 2 ϕ(y) ≤ 1 2 + 1 2 = 1, we see ϕ(x) = ϕ(y) = 1. For the elements of the unit sphere in X given by x and y (denoted by fx and fy), we have fx(ϕ) = fy(ϕ) = 1. Since fx = fy, the space X cannot be smooth at ϕ. Corollary 3.26. Let X be reflexive. The space X is smooth if and only if X is strictly convex; and X is strictly convex if and only if X is smooth. Remark 3.27. Note that there exist smooth spaces, whose dual spaces are not strictly convex. Similarly, there exist strictly convex spaces, whose dual spaces are not smooth. We repeat that smooth Banach spaces are the spaces whose norms have the Gâteaux derivative at any point of the unit sphere. Definition 3.28. The space X is called uniformly smooth if there exists the limit lim τ→0 x+τy − x τ uniformly for x,y ∈ ∂B(0,1). We end this chapter with the analogy of Corollary 3.26. Theorem 3.29. The space X is uniformly smooth if and only if X is uniformly convex; and X is uniformly convex if and only if X is uniformly smooth. We add that uniformly smooth spaces are reflexive. 48 4. Fixed point theorems In this chapter, we consider a real Banach space X. Definition 4.1. Let Y be a metric space. A point y is called a fixed point of a map f : D ⊆ Y → Y if f(y) = y. At first, we recall the most important fixed point theorem. Theorem 4.2 (Banach). If Y is a complete metric space and f : Y →Y is a contraction, then there exists just one fixed point of f. Remark 4.3. Theorem 4.2 is well-known, e.g., from the theory of ODEs. Definition 4.4. We say that a subset D of X has the fixed point property if any continuous map f : D → D has a fixed point. Remark 4.5. For example, for (R,|·|), it is seen that any closed interval [a,b] has the fixed point property. 4.1 Topological degree Let f : G → X be a map, where G is an open set in X. Our goal is to define the number deg(f,G, p), i.e., the so-called topological degree of f, which means “the number of solutions of the equation f(x) = p on G”. This number depends on f and on p continuously in a certain sense. Thus, for some perturbations of f, the number deg(f,G, p) has to be constant in a sufficiently small neighbourhood of p. Let us consider the Euclidean space Rn . Definition 4.6. To any (f,G, p), where G is from the system of all bounded open subsets of Rn , f : G → Rn is a continuous map, and p ∈ Rn f(∂G), we assign the integer deg(f,G, p) satisfying the following conditions: i) if f is the identity on G and p ∈ G, then deg(f,G, p) = 1; ii) if G1,G2 ⊆ G are open sets satisfying G1 ∩ G2 = /0 and p /∈ f G (G1 ∪G2) , then deg(f,G, p) = deg(f,G1, p)+deg(f,G2, p); iii) if H : [0,1]×G → Rn is a continuous map, f0(x) = H(0,x), f1(x) = H(1,x), and if H(t,x) = p for t ∈ [0,1], x ∈ ∂G, then deg(f0,G, p) = deg(f1,G, p); iv) if deg(f,G, p) = 0, then there exists x ∈ G such that f(x) = p. This map is called the topological degree in Rn . Remark 4.7. In Rn , the map from the previous definition exists and it is determined by the conditions i)–iv) uniquely. 49 Before Remark 4.9 mentioned below, we recall the notion of the so-called homo- topy. Definition 4.8. Let Z and Y be metric spaces and let f,g: Z → Y be continuous maps. We say that f is homotopic with g if there exists a continuous map H : [0,1] × Z → Y such that H(0,x) = f(x) and H(1,x) = g(x) for x ∈ Z. The map H is called the homotopy between f,g. Remark 4.9. Now, we comment the conditions i)–iv) from Definition 4.6. The condition i) says that the equation idx = p has one solution x = p. The condition ii) says that, if the equation f(x) = p has just n1 solutions on G1, just n2 solutions on G2, and no solution on G (G1 ∪ G2), then this equation has n1 + n2 solutions on G. The condition iii) expresses the invariance of the topological degree with respect to homotopies. The condition iv) says when the equation f(x) = p has solutions on G. Remark 4.10. Now, using a simple example, we explain why we cannot consider also p ∈ f(∂G) in Definition 4.6. Let us consider the function f(t) = t on G = (0,1) ⊆ R. If p ∈ (−∞,0) ∪ (1,∞), then deg(f,G, p) = 0, because the equation f(x) = p has no solution from the interval (0,1) for this p. Next, deg(f,G, p) = 1 for p ∈ (0,1). In any neighbourhood of p = 0 or p = 1, the topological degree deg(f,G, p) takes the both values 0 and 1. Hence, for p ∈ f(∂G) = {0,1}, deg(f,G, p) cannot be defined if the topological degree depends on p continuously. Remark 4.11. In spaces whose dimension is infinite, one can construct the theory of the topological degree as well. It is called the Leray–Schauder degree and it is introduced for maps of the type I −T, where T is, e.g., a completely continuous linear operator. Theorem 4.12. The topological degree in Rn has the following properties: 1. if f,g: G → Rn are continuous maps, f = g on ∂G, and p ∈ Rn f(∂G), then deg(f,G, p) = deg(g,G, p); 2. the map deg(f,G,−) is constant on any connected component of the open set Rn f (∂G). 4.2 Brouwer and 1. and 2. Schauder theorem Now, we use the topological degree in the Euclidean space Rn . Theorem 4.13 (Brouwer). The closed unit ball B[0,1] ⊆ Rn has the fixed point pro- perty. Proof. By contradiction, we consider a continuous map f : B[0,1] → B[0,1] with the property that f(x) = x for all x ∈ B[0,1]. The map H(t,x) = x−t f(x), x ∈ Rn , t ∈ [0,1], is a homotopy. We show that H(t,x) = 0 for t ∈ [0,1] and x = 1. For t = 1, it follows from the assumption. If t ∈ [0,1), then we have the inequality t f(x) ≤ t < 1 for x = 1. Therefore, x = t f(x). We denote g0(x) = H(0,x) = x, g1(x) = H(1,x) = x− f(x), x ∈ B(0,1). 50 From the condition iii) in Definition 4.6, we obtain that deg(g0,B(0,1),0) = deg(g1,B(0,1),0). But, g0 is the identity, which gives deg(g0,B(0,1),0) = 1. Thus, deg(g1,B(0,1),0) = 1 and the condition iv) from Definition 4.6 gives the existence of x ∈ B(0,1) for which g1(x) = 0. Of course, this is a contradiction. Remark 4.14. From Theorem 4.13, it follows that any compact convex subset of a Banach space with a finite dimension has the fixed point property. Naturally, we obtain the question, whether the Brouwer theorem, i.e., Theorem 4.13, is valid also for the closed unit balls in spaces whose dimension is infinite. But, it is enough to consider, e.g., the map h: x = {x1,x2, ...} → 1− x 2 ,x1,x2, ... , which does not have any fixed point on the closed unit ball of the space l2 . For spaces whose dimension is infinite, we have the following result. Theorem 4.15 (Schauder). Let K be a (non-empty) compact convex subset of X and let f : K → K be a continuous map. Then, there exists x ∈ K such that f(x) = x. Proof. We consider ε > 0. There exist x1,...,xn ∈ K such that K ⊆ n j=1 B(xj,ε). We define ϕj(x) = max 0,ε − x−xj , x ∈ K, j ∈ {1,...,n}. The functions ϕj are non-negative on K and the function n ∑ j=1 ϕj is positive on K. Thus, we can define the function ϕ on K by ϕ : x → n ∑ j=1 ϕj(x)xj n ∑ j=1 ϕj(x) −1 . Obviously, ϕ is a continuous function on K which maps K to the set Kε = conv{x1,...,xn} ⊆ K, where conv{x1,...,xn} is the convex hull of x1,...,xn. We have ϕ(x)−x ≤ ε, x ∈ K. 51 The composition ϕ ◦ f maps Kε into Kε . According to Theorem 4.13 (see Remark 4.14), it has a fixed point xε ∈ Kε . Since xε − f(xε ) ≤ xε −ϕ(f(xε )) + ϕ(f(xε ))− f(xε ) = ϕ(f(xε ))− f(xε ) ≤ ε, we have inf{ x− f(x) ; x ∈ K} = 0. Considering that f is a continuous map on a compact set, we know that there exists x ∈ K such that f(x) = x. Definition 4.16. A map f : D ⊆ X → X is called compact if it is continuous and if it maps any bounded subset of D into a set whose closure is a compact set. Remark 4.17. We repeat that a linear map f : X → X, which maps bounded sets into sets with compact closures, is continuous. Theorem 4.18 (Schauder). Let K be a closed bounded convex subset of X and let f : K → K be compact. Then, f has a fixed point. 52 5. Integration in Banach spaces In this chapter, we consider a real Banach space X. 5.1 Preliminaries and basic definitions At first, we recall basic definitions. Definition 5.1. A system S of subsets of a given set Ω is called a σ-algebra if a) Ω ∈ S; b) A ∈ S ⇒ Ω A ∈ S; c) An ∈ S,n ∈ N ⇒ ∞ n=1 An ∈ S. Then, (Ω,S) is called a measurable space. Definition 5.2. Let S be a system of subsets of a set Ω. A map µ : S → [0,+∞] is called a measure if: a) S is σ-algebra; b) µ(/0) = 0; c) for any sequence {An}∞ n=1 of pairwise disjoint sets from S, it holds µ ∞ n=1 An = ∞ ∑ n=1 µ(An). We say that a measure µ is finite if µ(Ω) < ∞, and it is probability if µ(Ω) = 1. A measure µ is called complete if the implication A,B ⊆ Ω, A ⊆ B, B ∈ S, µ(B) = 0 =⇒ A ∈ S is valid. Let a measurable space (Ω,S) be given with a measure µ, where µ is a probability complete measure. Definition 5.3. Let Y be an arbitrary set and let A ⊆ Y. We define the map χA : Y → R by χA(y) = 1 for y ∈ A; 0 for y /∈ A. Definition 5.4. A map f : Ω → X is called 53 • simple if there exist x1,...,xn ∈ X and E1,...,En ∈ S such that f ≡ n ∑ i=1 xiχEi ; • measurable if there exists a sequence { fn}∞ n=1 of simple maps such that lim n→∞ fn(ω) = f(ω) for all ω ∈ Ω up to a set whose measure is zero; • weakly measurable if ϕ ◦ f is a measurable function for all ϕ ∈ X . Remark 5.5. It is easy to show that f1 + f2 and λ f1 are measurable if f1, f2 are measurable and λ ∈ R (analogously, in the weakly measurable case). Theorem 5.6 (Pettis). A map f : Ω → X is measurable if and only if f is weakly measurable and there exists a set E ∈ S with µ(E) = 0 such that f (Ω E) is a separable subset of X. Especially, for separable Banach spaces, the notions of measurable and weakly measurable maps are same. Definition 5.7. If xn (for n ∈ N) are elements of the Banach space X, then we say that the series ∞ ∑ i=1 xi • converges if there exists lim n→∞ n ∑ i=1 xi; • converges absolutely if ∞ ∑ i=1 xi < +∞; • converges unconditionally to x ∈ X if ∞ ∑ i=1 xp(i) = x for all permutation p of N. Remark 5.8. Due to the completeness of X, any absolutely convergent series is also unconditionally convergent. If the dimension of X is finite, then any unconditionally convergent series converges absolutely as well. In spaces whose dimension is infinite, it is not valid. It suffices to consider xn = 0,...,0, 1 n ,0, ... ∈ c0, n ∈ N, where c0 is the space of all sequences {xn}∞ n=1 of real numbers which converge to zero with the norm x = max n∈N |xn|. 54 Definition 5.9. Let F : S → X. We say that F is an additive or σ-additive vector measure if F(/0) = 0 and F n En = ∑ n F(En) for all finite or countable sequence of disjoint sets En ∈ S, respectively. Remark 5.10. The convergence in Definition 5.9 is the convergence of series in X (in the case of the considered σ-additivity). This convergence is unconditional. Definition 5.11. We say that a vector measure F : S → X is absolutely continuous with respect to the measure µ if, for any ε > 0, there exists δ > 0 such that F(E) < ε if µ(E) < δ. Theorem 5.12 (Pettis). Let F be a σ-additive vector measure. Then, F is absolutely continuous with respect to the measure µ if and only if F(E) = 0 for µ(E) = 0. Definition 5.13. Let F : S → X be a vector measure. If |F(Ω)| = sup n ∑ k=1 F(Ak) ; Ak ∈ S are pairwise disjoint, n k=1 Ak = Ω < +∞, then we say that F has a bounded variation. 5.2 Bochner integral Definition 5.14. A map f : Ω → X is called integrable in the Bochner sense if f is measurable and if there exist simple maps fn for n ∈ N such that lim n→∞ Ω f − fn dµ = 0. One can show the following implication. If lim n→∞ B f − fn dµ = 0 = lim n→∞ B f −gn dµ for a measurable map f, simple maps fn,gn, n ∈ N, and B ∈ S, then lim n→∞ B fn dµ = lim n→∞ B gn dµ ∈ X, where we put B ϕ dµ = p ∑ i=1 xiµ (Ei ∩B) for a simple map ϕ ≡ p ∑ i=1 xiχEi , The previous implication guarantees the correctness of the following definition. 55 Definition 5.15. Let f be integrable in the Bochner sense, let B ∈ S, and let { fn}∞ n=1 be a sequence of simple maps satisfying lim n→∞ B f − fn dµ = 0. The limit lim n→∞ B fn dµ ∈ X is called the Bochner integral of f over B and it is denoted by B f dµ. Theorem 5.16 (Bochner). Let f : Ω → X be measurable. Then, f is integrable in the Bochner sense if and only if Ω f dµ < +∞. As L1 X (or L1 X (Ω,S,µ)), we denote the space of all maps which are integrable in the Bochner sense. Theorem 5.17. For f ∈ L1 X and E ∈ S, the inequality E f dµ ≤ E f dµ is valid. Theorem 5.18. Let us consider the identification of maps which are different on sets with zero measures. Then, L1 X is a Banach space with the norm f L1 X = Ω f dµ. Definition 5.19. The indefinite Bochner integral F : S → X is defined by F(E) = E f dµ, where E ∈ S. Theorem 5.20. Let a map f : Ω → X be integrable in the Bochner sense. The indefinite Bochner integral is a σ-additive vector measure, which is absolutely continuous with respect to µ. If En ∈ S are pairwise disjoint, then the series ∑ n F(En) converges absolutely. 56 5.3 Gelfand, Pettis, and Darboux integral Lemma 5.21. Let f : Ω → X. If ϕ ◦ f ∈ L1 R for all ϕ ∈ X , then, for any E ∈ S, there exists LE ∈ X such that LE(ϕ) = E ϕ ◦ f dµ for all ϕ ∈ X . Definition 5.22. Let f : Ω → X. If ϕ ◦ f ∈ L1 R for all ϕ ∈ X , then we say that f is weakly integrable. The element LE ∈ X , whose existence is guaranteed by Lemma 5.21, is called the Gelfand integral of f over E. Similarly as for the Bochner integral, one can introduce the indefinite Gelfand integral, which maps E ∈ S into LE ∈ X . In addition, if LE ∈ X for all E ∈ S, i.e., there exists pE ∈ X such that LE(ϕ) = ϕ(pE) for all ϕ ∈ X , then we say that f is integrable in the Pettis sense. The element pE ∈ X is called the Pettis integral of f over E. The indefinite Pettis integral maps E ∈ S into pE ∈ X. Remark 5.23. If X is reflexive, then the Gelfand integral and the Pettis integral are identical. Theorem 5.24. Let a map f : Ω → X be weakly integrable. Then, the following statements are equivalent: i) f is integrable in the Pettis sense; ii) the indefinite Gelfand integral of f is a σ-additive vector measure; iii) the indefinite Gelfand integral of f is absolutely continuous with respect to µ. Remark 5.25. If a map f is integrable in the Bochner sense, then f is integrable in the Pettis sense as well. More specifically, the relationship between the “strong” integral and the “weak” integral is described in the following theorem. Theorem 5.26. Let f : Ω → X be measurable and integrable in the Pettis sense. Then, f is integrable in the Bochner sense if and only if the indefinite Pettis integral of f is a vector measure having a bounded variation. Theorem 5.27. Let K be a compact subset of X and let µ be a probability complete measure on K. If f : K → X is continuous, then there exists the Pettis integral of f over K. Definition 5.28. Let a map f : [0,1] → X be given. Let D = {0 = x0 < x1 < ··· < xn = 1} be a partition of the interval [0,1] and let D(f,D) = n ∑ i=1 sup{ f(s)− f(t) ;s,t ∈ [xi−1,xi]}·(xi −xi−1). We say that f is integrable in the Darboux sense if, for all ε > 0, there exists δ > 0 such that D(f,D) < ε if the norm of the partition D is less than δ. 57 Theorem 5.29. A map f : [0,1] → X is integrable in the Darboux sense if and only if f is a bounded map which is continuous on [0,1] up to a set whose Lebesgue measure is zero. Theorem 5.30. Any map f : [0,1] → X integrable in the Darboux sense is integrable in the Bochner sense. 58 REFERENCES [1] A. N. Kolmogorov; S. V. Fomin: Základy teorie funkcí a funkcionální analýzy. Praha: SNTL – Nakladatelství technické literatury, 1975. [2] J. Lukeš: Úvod do funkcionální analýzy. Praha: Karolinum, 2005. [3] J. Lukeš: Zápisky z funkcionální analýzy. Praha: Karolinum, 2002. [4] J. Lukeš; J. Malý: Míra a integrál. Praha: Karolinum, 2002. [5] M. L. Einsiedler: Functional analysis, spectral theory, and applications. Cham: Springer, 2017. [6] D. R. Farenick: Fundamentals of functional analysis. Cham: Springer, 2016. [7] E. Kreyszig: Introductory functional analysis with applications. New York: Wiley, 1989. [8] J. T. Oden; L. Demkowicz: Applied functional analysis. Boca Raton: Chapman and Hall/CRC, 2018. [9] A. H. Siddiqi: Functional analysis and applications, Singapore: Springer, 2018.