In this appendix, we present definitions of basic terminology used in the book for the reader’s convenience. For a given set W, \(x\in W\) means that x is an element of W.

5.1 Convergence

  1. (1)

    Let M be a set. A real-valued function d defined on \(M\times M\) is said to be a metric if

    1. (i)

      \(d(x,y)=0\) if and only if \(x=y\) for \(x,y\in M\);

    2. (ii)

      (symmetry) \(d(x,y)=d(y,x)\) for all \(x,y\in M\);

    3. (iii)

      (triangle inequality) \(d(x,y)\leq d(x,z)+d(z,y)\) for all \(x,y,z\in M\).

    Here, \(X_1\times X_2\) denotes the Cartesian product of two sets \(X_1\) and \(X_2\) defined by

    $$\displaystyle \begin{aligned} X_1 \times X_2 := \left\{ (x_1,x_2) \bigm| x_i \in X_i\ \text{for}\ i=1,2 \right\}. \end{aligned}$$

    The set M equipped with a metric d is called a metric space and denoted by \((M,d)\) if one needs to clarify the metric. Let W be a product of metric spaces of \((M_i,d_i)\) (\(i=1,\ldots ,m\)), i.e.,

    $$\displaystyle \begin{aligned} W &= \prod_{i=1}^m M_i = M_1 \times\cdots\times M_m \\ &:= \left\{ (x_1,\ldots,x_m) \bigm| x_i \in M_i \ \text{for}\ i =1, \ldots, m \right\}. \end{aligned} $$

    This W is metrizable, for example, with a metric

    $$\displaystyle \begin{aligned} d(x,y) = \left( \sum_{i=1}^m d_i(x_i, y_i)^2\right)^{1/2} \end{aligned}$$

    for \(x=(x_1,\ldots ,x_m)\), \(y=(y_1,\ldots ,y_m)\in W\). If \(M_i\) is independent of i, i.e., \(M_i=M\), then we simply write W as \(M^m\).

    A subset A of M is said to be open if for any \(x\in A\) there is \(\varepsilon >0\) such that the ball\(B_\varepsilon (x)=\left \{y\in M\bigm | d(y,x)\leq \varepsilon \right \}\) is included in A. If the complement\(A^c\) is open, then A is said to be closed. The complement \(A^c\) is defined by

    $$\displaystyle \begin{aligned} A^c = M \backslash A := \left\{ x \in M \bigm| x \not\in A \right\}. \end{aligned}$$

    For a set A, the smallest closed set including A is called the closure of A and denoted by \(\overline {A}\). Similarly, the largest open set included in A is called the interior of A and denoted by \(\operatorname {int}A\) or simply by \(\mathring {A}\). By definition, \(A=\overline {A}\) if and only if A is closed, and \(A=\mathring {A}\) if and only if A is open. The set \(\overline {A}\backslash \mathring {A}\) is called the boundary of A and denoted by \(\partial A\). For a subset B of a set A, we say that B is dense in A if \(\overline {B}=A\). A set A in M is bounded if there is \(x_0\in M\) and \(R>0\) such that A is included in \(B_R(x_0)\). For a mappingf from a set S to M (i.e., an M-valued function defined on S), f is said to be bounded if its image\(f(S)\) is bounded in M, where

    $$\displaystyle \begin{aligned} f(S) = \left\{ f(x) \bigm| x \in S \right\}. \end{aligned}$$
  2. (2)

    Let V  be a real vector space (a vector space over the field \(\mathbf {R}\)). A nonnegative function \(\|\cdot \|\) on V  is said to be a norm if

    1. (i)

      \(\|x\|=0\) if and only if \(x=0\) for \(x\in V\);

    2. (ii)

      \(\|cx\|=|c|\|x\|\) for all \(x\in V\) and all \(c\in \mathbf {R}\);

    3. (iii)

      (triangle inequality) \(\|x+y\| \leq \|x\|+\|y\|\) for all \(x,y\in V\).

    The vector space V  equipped with a norm \(\|\cdot \|\) is called a normed vector space and denoted by \(\left (V,\|\cdot \|\right )\) if one needs to clarify the norm. By definition,

    $$\displaystyle \begin{aligned} d(x,y) = \| x-y \| \end{aligned}$$

    is a metric. A normed vector space is regarded as a metric space with the foregoing metric.

  3. (3)

    Let \(\{z_j\}^\infty _{j=1}\) be a sequence in a metric space \((M,d)\). We say that \(\{z_j\}^\infty _{j=1}\)converges to \(z\in M\) if for any \(\varepsilon >0\) there exists a natural number \(n=n(\varepsilon )\) such that \(j\geq n(\varepsilon )\) implies \(d(z,z_j)<\varepsilon \). In other words,

    $$\displaystyle \begin{aligned} \lim_{j \to 0} d(z,z_j) = 0. \end{aligned}$$

    We simply write \(z_j\to z\) as \(j\to \infty \), or \(\lim _{j\to \infty }z_j=z\). If \(\{z_j\}^\infty _{j=1}\) converges to some element, we say that \(\{z_j\}^\infty _{j=1}\) is a convergent sequence.

  4. (4)

    Let f be a mapping from a metric space \((M_1,d_1)\) to another metric space \((M_2,d_2)\). We say that \(f(y)\)converges to \(a\in M_2\) as y tends to x if for any \(\varepsilon >0\) there exists \(\delta =\delta (\varepsilon )>0\) such that

    $$\displaystyle \begin{aligned} d_2 \left(f(y), a\right) < \varepsilon \quad \text{if}\quad d_1(y,x) < \delta. \end{aligned}$$

    We simply write \(f(y)\to a \) as \(y\to x\) or \(\lim _{y\to x}f(y)=a\). If

    $$\displaystyle \begin{aligned} \lim_{y \to x} f(y) = f(x), \end{aligned}$$

    then f is said to be continuous at \(x\in M_1\). If f is continuous at all \(x\in M_1\), then f is said to be continuous on \(M_1\) (with values in \(M_2\)). The space of all continuous functions on \(M_1\) with values in \(M_2\) is denoted by \(C(M_1, M_2)\).

  5. (5)

    Let \(\{z_j\}^\infty _{j=1}\) be a sequence in a metric space \((M,d)\). We say that \(\{z_j\}^\infty _{j=1}\) is a Cauchy sequence if for any \(\varepsilon >0\) there exists a natural number \(n=n(\varepsilon )\) such that \(j,k\geq n(\varepsilon )\) implies \(d(z_j,z_k)<\varepsilon \). It is easy to see that a convergent sequence is always a Cauchy sequence, but the converse may not hold. We say that the metric space \((M,d)\) is complete if any Cauchy sequence is a convergent sequence.

  6. (6)

    Let \(\left (V,\|\cdot \|\right )\) be a normed vector space. We say that V  is a Banach space if it is complete as a metric space. The norm \(\|\cdot \|\) is often written as \(\|\cdot \|{ }_V\) to distinguish it from other norms if we use several norms. We simply write \(z_j\to z\) in V  (as \(j\to \infty \)) if \(\lim _{j\to \infty }\|z_j-z\|{ }_V=0\) and \(z\in V\) for a sequence \(\{z_j\}_{j=1}^\infty \). We often say that \(z_j\) converges to zstrongly in V  (as \(j\to \infty \)) to distinguish this convergence from other weaker convergences discussed later.

  7. (7)

    Let V  be a real vector space. A real-valued function \(\langle \cdot ,\cdot \rangle \) defined on \(V\times V\) is said to be an inner product if

    1. (i)

      \(\langle x,x \rangle \geq 0\) for all \(x\in V\);

    2. (ii)

      \(\langle x,x \rangle = 0\) if and only if \(x=0\);

    3. (iii)

      (symmetry) \(\langle x,y \rangle =\langle y,x \rangle \) for all \(x,y \in V\);

    4. (iv)

      (linearity) \(\langle c_1 x_1 + c_2 x_2, y \rangle =c_1\langle x_1, y \rangle + c_2\langle x_2, y \rangle \) for all \(x_1, x_2, y \in V\), \(c_1, c_2 \in \mathbf {R}\).

    By definition, it is easy to see that

    $$\displaystyle \begin{aligned} \|z\| = \langle z, z \rangle^{1/2} \end{aligned}$$

    is a norm. The space with an inner product is regarded as a normed vector space with the foregoing norm. If this space is complete as a metric space, we say that V  is a Hilbert space. The Euclidean space \({\mathbf {R}}^N\) is a finite-dimensional Hilbert space equipped with a standard inner product. It turns out that any finite-dimensional Hilbert space is “isomorphic” to \({\mathbf {R}}^N\). Of course, a Hilbert space is an example of a Banach space.

  8. (8)

    Let V  be a Banach space equipped with norm \(\|\cdot \|\). Let \(V^*\) denote the totality of all continuous linear function(al)s on V  with values in \(\mathbf {R}\). (By the Hahn–Banach theorem, the vector space \(V^*\) has at least one dimension. Incidentally, Mazur’s theorem in the proof of Lemma 1.19 in Sect. 1.2.3 is another application of the Hahn–Banach theorem.)

    The space \(V^*\) is called the dual space of V . Let \(\{z_j\}^\infty _{j=1}\) be a sequence in \(V^*\). We say that \(\{z_j\}^\infty _{j=1}\) converges to \(z\in V^* \ *\)-weakly if

    $$\displaystyle \begin{aligned} \lim_{j\to\infty} z_j(x) = z(x) \end{aligned}$$

    for any \(x\in V\). We often write \(z_j\overset {*}{\rightharpoonup } z\) in \(V^*\) as \(j\to \infty \). Such a sequence \(\{z_j\}^\infty _{j=1}\) is called a \(*\)-weak convergent sequence. The dual space \(V^*\) is equipped with the norm

    $$\displaystyle \begin{aligned} \|z\|{}_{V^*} := \sup \left\{ z(x) \bigm| \|x\| = 1,\ x \in V \right\} = \sup_{\|x\|=1}z(x). \end{aligned}$$

    The space \(V^*\) is also a Banach space with this norm. Here, for a subset A in \(\mathbf {R}\), by \(a=\sup A\) we mean that a is the smallest real member that satisfies \(a\geq x\) for any \(a\in A\). In other words, it is the least upper bound of A. The notation \(\sup \) is the abbreviation of the supremum. Similarly, \(\inf A\) denotes the greatest lower bound of A, and it is the abbreviation of the infimum. If \(\sup A=a\) with \(a\in A\), we write \(\max A\) instead of \(\sup A\). The same convention applies to \(\inf \) and \(\min \).

    Since \(V^*\) is a Banach space, there is a notion of convergence in the metric defined by the norm. To distinguish this convergence from \(*\)-weak convergence, we say that \(\{z_j\}^\infty _{j=1}\) converges to zstrongly in \(V^*\) if

    $$\displaystyle \begin{aligned} \lim_{j\to\infty} \|z_j-z\|{}_{V^*} = 0, \end{aligned}$$

    and it is simply written \(z_j\to z\) in \(V^*\) as \(j\to \infty \). By definition, \(z_j\to z\) implies \(z_j \overset {*}{\rightharpoonup } z\), but the converse may not hold.

  9. (9)

    Let A be a subset of a metric space M. The set A is said to be (sequentially) relatively compact if any sequence \(\{z_j\}^\infty _{j=1}\) in A has a convergent subsequence in M. If, moreover, A is closed, we simply say that A is compact. When A is compact, it is always bounded. When A is a subset of \({\mathbf {R}}^N\), it is well known as the Bolzano–Weierstrass theorem that A is compact if and only if A is bounded and closed. However, if A is a subset of a Banach space V , such an equivalence holds if and only if V  is of finite dimension. In other words, a bounded sequence of an infinite-dimensional Banach space may not have a (strongly) convergent subsequence.

    There is a compactness theorem (Banach–Alaoglu theorem) that says if \(\{z_j\}^\infty _{j=1}\) in a dual Banach space \(V^*\) is bounded, i.e.,

    $$\displaystyle \begin{aligned} \sup_{j\geq1}\|z_j\|{}_{V^*} < \infty, \end{aligned}$$

    then it has a \(*\)-weak convergent subsequence (Exercise 1.9).

  10. (10)

    Let V  be a Banach space and \(V^*\) denote its dual space. Let \(\{x_k\}^\infty _{k=1}\) be a sequence in V . We say that \(\{x_k\}^\infty _{k=1}\) converges to \(x\in V\)weakly if

    $$\displaystyle \begin{aligned} \lim_{k\to\infty} z(x_k) = z(x) \end{aligned}$$

    for all \(z\in V^*\). We often write \(x_k \rightharpoonup x\) in V  as \(k\to \infty \). Such a sequence is called a weak convergent sequence.

    If a Banach space W is a dual space of some Banach space V , say, \(W=V^*\), there are two notions, weak convergence and \(*\)-weak convergence. Let \(\{z_j\}_{j=1}^\infty \) be a sequence in W. By definition, \(z_j \overset {*}{\rightharpoonup } z\) (in W as \(j\to \infty \)) means that \(\lim _{j\to \infty }z_j(x)=z(x)\) for all \(x\in V\) while \(z_j \rightharpoonup z\) (in W as \(j\to \infty \)) means that \(\lim _{j\to \infty }y(z_j)=y(z)\) for all \(y\in W^*=(V^*)^*\).

    The space V  can be continuously embedded in \(V^{**}=(V^*)^*\). However, V  may not be equal to \(V^{**}\). Thus, weak convergence is stronger than \(*\)-weak convergence. If \(V=V^{**}\), then both notions are the same. The space V  is called reflexive if \(V=V^{**}\).

  11. (11)

    If V  is a Hilbert space, it is reflexive. More precisely, the mapping \(x\in V\) to \(z\in V^*\) defined by

    $$\displaystyle \begin{aligned} z(y) = \langle x,y \rangle, \quad y \in V \end{aligned}$$

    is a linear isomorphism from V  to \(V^*\), which is also norm preserving, i.e., \(\|z\|{ }_{V^*}=\|x\|\). This result is known as the Riesz–Fréchet theorem. Thus, the notions of weak convergence and \(*\)-weak convergence are the same.

  12. (12)

    Let f be a real-valued function in a metric space M. We say that f is lower semicontinuous at \(x\in M\) if

    $$\displaystyle \begin{aligned} f(x) \leq \liminf_{y\to x} f(y) := \lim_{\delta\downarrow 0}\inf \left\{ f(y) \bigm| d(y,x)<\delta \right\}, \end{aligned}$$

    where \(\lim _{\delta \downarrow 0}\) denotes the limit as \(\delta \to 0\) but restricted to \(\delta >0\). Even if f is allowed to take \(+\infty \), the definition of the lower semicontinuity will still be valid. If f is lower semicontinuous for all \(x\in M\), we simply say that f is lower semicontinuous on M. If \(-f\) is lower semicontinuous, we say that f is upper semicontinuous.

  13. (13)

    Let \(f=f(t)\) be a function of one variable in an interval I in \(\mathbf {R}\) with values in a Banach space V . We say that f is right differentiable at \(t_0\in I\) if there is \(v\in V\) such that

    $$\displaystyle \begin{aligned} \lim_{h\downarrow 0} \left\| f(t_0 + h) - f(t_0) - vh \right\| \bigm/h = 0 \end{aligned}$$

    provided that \(t_0+h\in I\) for sufficiently small \(h>0\). Such v is uniquely determined if it exists and is denoted by

    $$\displaystyle \begin{aligned} v = \frac{\mathrm{d}^+ f}{\mathrm{d}t} (t_0). \end{aligned}$$

    This quantity is called the right differential of f at \(t_0\). The function \(t\mapsto \frac {\mathrm {d}^+f}{\mathrm {d}t}(t)\) is called the right derivative of f. The left differentiability is defined in a symmetric way by replacing \(h\downarrow 0\) with \(h\uparrow 0\). Even if both right and left differentials exist, they may be different. For example, consider \(f(t)=|t|\) at \(t_0=0\). The right differential at zero is 1, while the left differential at zero is \(-1\). If the right and left differentials agree with each other at \(t=t_0\), we say that f is differentiable at \(t=t_0\), and its value is denoted by \(\frac {\mathrm {d}f}{\mathrm {d}t}(t_0)\). The function \(t\mapsto \frac {\mathrm {d}f}{\mathrm {d}t}(t)\) is called the derivative of f. If f depends on other variables, we write \(\partial f/\partial t\) instead of \(\mathrm {d}f/\mathrm {d}t\) and call the partial derivative of f with respect to t.

5.2 Measures and Integrals

  1. (1)

    For a set M, let \(2^M\) denote the family of all subsets of H. We say that a function \(\mu \) defined on \(2^M\) with values in \([0,\infty ]\) is an (outer) measure if

    1. (i)

      \(\mu (\emptyset )=0\);

    2. (ii)

      (countable subadditivity) \(\mu (A)\leq \sum ^\infty _{j=1} \mu (A_j)\) if a countable family \(\{A_j\}^\infty _{j=1}\)coversA, where \(A_j,A\in 2^M\). In other words, A is included in a union of \(\{A_j\}^\infty _{j=1}\), i.e., a point of A must be an element of some \(A_j\).

    Here, \(\emptyset \) denotes the empty set.

  2. (2)

    A set \(A\in 2^M\) is said to be \(\mu \)-measurable if

    $$\displaystyle \begin{aligned} \mu(S \cap A^c) + \mu(S \cap A) = \mu(S) \end{aligned}$$

    for any \(S\in 2^M\). Let \(M_0\) be a metric space. A mapping f from M to \(M_0\) is said to be \(\mu \)-measurable if the preimage\(f^{-1}(U)\) of an open set U of \(M_0\) is \(\mu \)-measurable. Here,

    $$\displaystyle \begin{aligned} f^{-1}(U) := \left\{ x \in M \bigm| f(x) \in U \right\}. \end{aligned}$$

    A set A with \(\mu (A)=0\) is called a \(\mu \)-measure zero set. If a statement \(P(x)\) for \(x\in M\) holds for \(x\in M\backslash A\) with \(\mu (A)=0\), we say that \(P(x)\) holds for \(\mu \)-almost every\(x\in M\) or shortly a.e. \(x\in M\). In other words, P holds in M outside a \(\mu \)-measure zero set. In this case, we simply say that P holds almost everywhere in M.

    Let \(\mathcal {M}\) be the set of all \(\mu \)-measurable sets. If we restrict \(\mu \) just to \(\mathcal {M}\), i.e., \(\overline {\mu }=\left .\mu \right |{ }_{\mathcal {M}}\), then \(\overline {\mu }\) becomes a measure on \(\mathcal {M}\). Since in this book we consider \(\mu (A)\) for a \(\mu \)-measurable set A, we often say simply a measure instead of an outer measure.

  3. (3)

    Let A be a subset of \({\mathbf {R}}^N\). Let \(\mathcal {C}\) be a family of closed cubes in \({\mathbf {R}}^N\) whose faces are orthogonal to the \(x_i\)-axis for some \(i=1,\ldots ,N\). In other words, \(C\in \mathcal {C}\) means

    $$\displaystyle \begin{aligned} C = \left\{ (x_1,\ldots,x_N) \in {\mathbf{R}}^N \bigm| a_i \leq x_i \leq a_i+\ell\ (i=1,\ldots,n) \right\} \end{aligned}$$

    for some \(a_i\), \(\ell \in \mathbf {R}\). Let \(|C|\) denote its volume, i.e., \(|C|=\ell ^N\). We set

    $$\displaystyle \begin{aligned} \mathcal{L}^N(A) = \inf \left\{ \sum^\infty_{j=1}|C_j| \Biggm| \{C_j\}^\infty_{j=1}\ \text{covers}\ A \ \text{with}\ C_j \in \mathcal{C} \right\}. \end{aligned}$$

    It turns out that \(\mathcal {L}^N(C)=|C|\); it is nontrivial to prove \(\mathcal {L}^N(C)\geq |C|\). It is easy to see that \(\mathcal {L}^N\) is an (outer) measure in \({\mathbf {R}}^N\). This measure is called the Lebesgue measure in \({\mathbf {R}}^N\). It can be regarded as a measure in the flat torus \({\mathbf {T}}^N=\prod ^N_{i=1}(\mathbf {R}/\omega _i\mathbf {Z})\). For a subset A of \({\mathbf {T}}^N\), we regard this set as a subset \(A_0\) of the fundamental domain (i.e., the periodic cell\([0,\omega _1)\times \cdots \times [0,\omega _N)\)). The Lebesgue measure of A is defined by \(\mathcal {L}^N(A)=\mathcal {L}^N(A_0)\). Evidently, \(\mathcal {L}^N({\mathbf {T}}^N)=\omega _1\cdots \omega _N\), which is denoted by \(|{\mathbf {T}}^N|\) in the proof of Lemma 1.21 in Sect. 1.2.5.

  4. (4)

    In this book, we only use the Lebesgue measure. We simply say measurable when a mapping or a set is \(\mathcal {L}^N\)-measurable. Instead of writing \(\mathcal {L}^N\)-a.e., we simply write a.e. Let \(\Omega \) be a measurable set in \({\mathbf {T}}^N\) or \({\mathbf {R}}^N\), for example, \(\Omega ={\mathbf {T}}^N\). Let f be a measurable function on \(\Omega \) with values in a Banach space V . Then one is able to define its integral over \(\Omega \). When \(V={\mathbf {R}}^m\), this integral is called the Lebesgue integral. In general, it is called the Bochner integral of f over \(\Omega \). Its value is denoted by \(\int _\Omega f\,\mathrm {d}\mathcal {L}^N\) or, simply, \(\int _\Omega f\,\mathrm {d}x\); See, for example, [90, Chapter V, Section 5]. If \(\Omega ={\mathbf {T}}^N\) and f is continuous, this agrees with the more conventional Riemann integral. For \(p\in [1,\infty )\) and a general Banach space V , let \(\tilde {L}^p(\Omega ,V)\) denote the space of all measurable functions f with values in V  such that

    $$\displaystyle \begin{aligned} \|f\|{}_p = \left( \int_\Omega \left\|f(x)\right\|{}^p \mathrm{d}x \right)^{1/p} \end{aligned}$$

    is finite. If \(\|f\|{ }_p\) is finite, we say that f is pth integrable. If \(p=1\), we simply say f is integrable. If \(f\in \tilde {L}^1(\Omega ,V)\), we say that f is integrable in \(\Omega \). We identify two functions \(f,g\in \tilde {L}^p(\Omega ,V)\) if \(f=g\) a.e. and define \(L^p(\Omega ,V)\) from \(\tilde {L}^p(\Omega ,V)\) by this identification. It is a fundamental result that \(L^p(\Omega ,V)\) is a Banach space equipped with the norm \(\|\cdot \|{ }_p\). When \(V=\mathbf {R}\), we simply write \(L^p(\Omega )\) instead of \(L^p(\Omega ,V)\). The case \(p=\infty \) should be handled separately. For a general Banach space V , let \(\tilde {L}^\infty (\Omega ,V)\) denote the space of all measurable functions f with values in V  such that

    $$\displaystyle \begin{aligned} \|f\|{}_\infty = \inf \left\{ \alpha \Bigm| \mathcal{L}^N \left( \left\{ x \in \Omega \bigm| \left\| f(x) \right\|{}_V > \alpha \right\} \right) = 0 \right\} \end{aligned}$$

    is finite. By the same identification, the space \(L^\infty (\Omega ,V)\) can be defined. This space \(L^\infty (\Omega ,V)\) is again a Banach space. Key theorems in the theory of Lebesgue integrals used in this book include the Lebesgue dominated convergence theorem and Fubini’s theorem. Here, we give a version of the dominated convergence theorem.

Theorem 5.1

Let V  be a Banach space. Let \(\{f_m\}_{m=1}^\infty \) be a sequence in \(L^1(\Omega ,V)\) . Assume that there is a nonnegative function \(\varphi \in L^1(\Omega )\) independent of m such that \(\left \|f_m(x)\right \|{ }_V\leq \varphi (x)\) for a.e. \(x\in \Omega \) . If \(\lim _{m\to \infty }f_m(x)=f(x)\) for a.e. \(x\in \Omega \) , then

$$\displaystyle \begin{aligned} \lim_{m\to\infty} \int_\Omega f_m(x)\, \mathrm{d}x = \int f(x)\,\mathrm{d}x. \end{aligned}$$

In other words,\(\lim _{m\to \infty } \left \|\int _\Omega f_m\, \mathrm {d}x - \int _\Omega f(x)\,\mathrm {d}x\right \|{ }_V=0\).

  • Usually, V  is taken as \(\mathbf {R}\) or \({\mathbf {R}}^N\), but it is easy to extend to this setting. For basic properties of the Lebesgue measure and integrals, see for example a classical book of Folland [42]. We take this opportunity to clarify \(*\)-weak convergence in \(L^p\) space. A basic fact is that \(\left (L^p(\Omega )\right )^*=L^{p'}(\Omega )\) for \(1\leq p<\infty \), where \(1/p+1/p'=1\). Note that \(p=\infty \) is excluded, but \((L^1)^*=L^\infty \). Since \(L^p\) is reflexive for \(1<p<\infty \), weak convergence and \(*\)-weak convergence agree with each other. Let us write a \(*\)-weak convergence in \(L^\infty \) explicitly. A sequence \(\{f_m\}\) in \(L^\infty (\Omega )\)\(*\)-weakly converges to \(f\in L^\infty (\Omega )\) as \(m\to \infty \) if and only if

    $$\displaystyle \begin{aligned} \lim_{m\to\infty} \int_\Omega f_m\varphi\, \mathrm{d}x = \int_\Omega f\varphi\, \mathrm{d}x \end{aligned}$$

    for all \(\varphi \in L^1(\Omega \)). For detailed properties of \(L^p\) spaces, see, for example, [19, Chapter 4].

    For a Banach space V -valued \(L^p\) function, we also consider its dual space. That is, we have

    $$\displaystyle \begin{aligned} \left(L^p(\Omega,V)\right)^* = L^{p'}(\Omega,V^*) \quad \text{for}\quad 1 \leq p < \infty \end{aligned}$$

    with \(1/p+1/p'=1\). (This duality—at least for reflexive V —can be proved along the same line as in [19, Chapter 4], where V  is assumed to be \(\mathbf {R}\). For a general Banach space V , see, for example, [35, Chapter IV].) We consider \(*\)-weak convergence in \(L^\infty (\Omega ,V)\) with \(V=L^q(U)\), \(1<q\leq \infty \), where \(\Omega \) is an open interval \((0,T)\) and U is an open set in \({\mathbf {T}}^N\) or \({\mathbf {R}}^N\) since this case is explicitly used in Chap. 2. A sequence \(\{f_m\}\) in \(L^\infty \left (\Omega ,L^q(U)\right )\)\(*\)-weakly converges to \(f\in L^\infty \left (\Omega ,L^q(U)\right )\) as \(m\to \infty \) if and only if

    $$\displaystyle \begin{aligned} \lim_{m\to\infty}\int_0^T \!\!\!\! \int_U f_m(x,t)\varphi(x,t)\mathrm{d}x\, \mathrm{d}t = \int_0^T \!\!\!\! \int_U f(x,t)\varphi(x,t)\mathrm{d}x\, \mathrm{d}t \end{aligned}$$

    for \(\varphi \in L^1\left (\Omega ,L^{q'}(U)\right )\). (Note that the space \(L^p\left (\Omega , L^q(U)\right )\) is identified with the space of all measurable functions \(\varphi \) on \(\Omega \times U\) such that \(\int _0^T \|\varphi \|{ }_{L^q(U)}^p(t)\, \mathrm {d}t<\infty \) or \(\int _0^T\left ( \int _U \left |\varphi (x,t)\right |{ }^q \mathrm {d}x\right )^{p/q}\, \mathrm {d}t<\infty \) for \(1\leq p,q<\infty \).)

  1. (5)

    Besides the basic properties of the Lebesgue integrals, we frequently use a few estimates involving \(L^p\)-norms. These properties are by now standard and found in many books, including [19]. For example, we frequently use the Hölder inequality

    $$\displaystyle \begin{aligned} \| fg \|{}_p \leq \|f\|{}_r \|g\|{}_q \end{aligned}$$

    with \(1/p=1/r+1/q\) for \(f\in L^r(\Omega ), g\in L^q(\Omega )\), where \(p,q,r\in [1,\infty ]\). Here, we interpret \(1/\infty =0\). In the case \(p=1\), \(r=q=2\), this inequality is called the Schwarz inequality. As an application, we have Young’s inequality for a convolution

    $$\displaystyle \begin{aligned} \|f*g\|{}_p \leq \|f\|{}_q \|g\|{}_r \end{aligned}$$

    for \(f\in L^q({\mathbf {R}}^N)\), \(g\in L^r({\mathbf {R}}^N)\) with \(1/p=1/q+1/r-1\) and \(p,q,r\in [1,\infty ]\); see, for example, [45, Chapter 4]. In this book, we use this inequality when \({\mathbf {R}}^N\) is replaced by \({\mathbf {T}}^N\).

  2. (6)

    In analysis, we often need an approximation of a function by smooth functions. We only recall an elementary fact. The space \(C^\infty _c(\Omega )\) is dense in \(L^p(\Omega )\) for \(p\in [1,\infty )\); see, for example, [19, Corollary 4.23]. However, it is not dense in \(L^\infty (\Omega )\).