Abstract:
In $L_2(\mathbb{R}^d;\mathbb{C}^n)$, we consider a selfadjoint strongly elliptic second-order differential operator ${\mathcal A}_\varepsilon$. It is assumed that the coefficients of ${\mathcal A}_\varepsilon$ are periodic and depend on ${\mathbf x}/\varepsilon$, where $\varepsilon>0$. We study the behaviour of the operator exponential $e^{-i{\mathcal A}_\varepsilon\tau}$ for small $\varepsilon$ and $\tau \in \mathbb{R}$. The results are applied to the homogenization of solutions of the Cauchy problem for the Schrödinger-type equation $i\partial_\tau{\mathbf u}_\varepsilon({\mathbf x},\tau)=({\mathcal A}_\varepsilon{\mathbf u}_\varepsilon)({\mathbf x},\tau)$ with initial data from a special class. For fixed $\tau$, as $\varepsilon \to 0$, the solution converges in $L_2(\mathbb{R}^d;\mathbb{C}^n)$ to the solution of the homogenized problem; the error is of the order $O(\varepsilon)$. For fixed $\tau$ we obtain an approximation of the solution ${\mathbf u}_\varepsilon(\,\cdot\,,\tau)$ in the $L_2(\mathbb{R}^d;\mathbb{C}^n)$-norm with error $O(\varepsilon^2)$, and also an approximation of the solution in the $H^1(\mathbb{R}^d;\mathbb{C}^n)$-norm with error $O(\varepsilon)$. In these approximations correctors are taken into account. The dependence of errors on the parameter $\tau$ is traced.
Bibliography: 113 items.
This paper concerns homogenization theory for periodic differential operators. An extensive literature is devoted to homogenization problems; first of all, we mention the monographs [1]–[3]. One of the methods of study of homogenization problems in $\mathbb{R}^d$ is the spectral method based on the Floquet–Bloch theory (see, for instance, [1], Chap. 4, [3], Chap. 2, [4], [5], and [6]).
0.1. The class of operators
We consider selfadjoint second-order differential operators acting on $L_2(\mathbb{R}^d;\mathbb{C}^n)$ and admitting a factorization of the form
Here $b({\mathbf D})=\sum_{l=1}^d b_l D_l$ is the first-order $m \times n$ matrix differential operator such that $m \geqslant n$ and the symbol $b(\boldsymbol{\xi})=\sum_{l=1}^d b_l \xi_l$ has the maximum rank. The matrix-valued functions $g({\mathbf x})$ (of size $m\times m$) and $f({\mathbf x})$ (of size $n\times n$) are periodic with respect to some lattice $\Gamma$; $g({\mathbf x})$ is positive definite and bounded; $f,f^{-1} \in L_\infty$. It is convenient to start with the study of the simpler class of operators given by
Many operators of mathematical physics can be written in the form (0.1) or (0.2); see [7] and [9], Chap. 4. The simplest example is the acoustics operator
Now we introduce the small parameter $\varepsilon>0$. For any $\Gamma$-periodic function $\varphi({\mathbf x})$ we set $\varphi^\varepsilon({\mathbf x}):=\varphi(\varepsilon^{-1}{\mathbf x})$. Consider the operators
0.2. Operator error estimates for second-order elliptic and parabolic equations in $\mathbb{R}^d$
In the series of papers [7]–[10] by Birman and Suslina an operator-theoretic approach to homogenization problems in $\mathbb{R}^d$ (a version of the spectral method) was proposed and developed. This approach is based on the scaling transformation, the Floquet–Bloch theory, and analytic perturbation theory.
Let us discuss results for the simpler operator (0.4). In [7] it was proved that
Here $\widehat{\mathcal{A}}^{\,0}=b({\mathbf D})^* g^0 b({\mathbf D})$ is the effective operator with constant effective matrix $g^0$. Approximations for the resolvent $(\widehat{\mathcal{A}}_\varepsilon +I)^{-1}$ in the $(L_2 \to L_2)$-norm with error term $O(\varepsilon^2)$ and in the $(L_2 \to H^1)$-norm with error term $O(\varepsilon)$ (with correctors taken into account) were obtained in [8], [9], and [10], respectively.
The operator-theoretic approach was applied to homogenization of parabolic problems in [11]–[16]. In [11] and [12] it was proved that
Approximations for the semigroup $e^{-\tau \widehat{\mathcal{A}}_\varepsilon}$ in the $(L_2 \to L_2)$-norm with error $O(\varepsilon^2)$ and in the $(L_2 \to H^1)$-norm with error $O(\varepsilon)$ (with correctors taken into account) were obtained in [13] and [14], respectively. Even more accurate approximations for the resolvent and the semigroup of the operator $\widehat{\mathcal{A}}_\varepsilon$ were found in [15] and [16].
The operator-theoretic approach was also applied to the more general class of operators $\widehat{\mathcal B}_\varepsilon$ with principal part $\widehat{\mathcal A}_\varepsilon$ and lower-order terms: the resolvent of such an operator was studied in [17] and [18] and the semigroup in [19] and [20].
Estimates of the form (0.5) and (0.6) are called operator error estimates in homogenization theory. They are order-sharp. A different approach to operator error estimates (the so-called shift method) was proposed by Zhikov and Pastukhova; see [21]–[23] and also the survey [24] and the references there.
0.3. Operator error estimates for Schrödinger-type equations and hyperbolic equations
The situation with homogenization of non-stationary Schrödinger-type equations and hyperbolic equations differs from the case of elliptic and parabolic problems. The operator-theoretic approach was applied to non-stationary problems in [25]. Again, let us dwell on results for the simpler operator (0.4). In operator terms, we are talking about approximations of the operators $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ and $\cos(\tau \widehat{\mathcal{A}}_\varepsilon^{\,1/2})$ (where $\tau \in \mathbb{R}$) for small $\varepsilon$. It has turned out that it is impossible to approximate these operators in the $(L_2 \to L_2)$-norm, and therefore the type of the operator norm should be changed. In [25] it was proved that
For the operator $\widehat{\mathcal{A}}_\varepsilon^{-1/2} \sin(\tau \widehat{\mathcal{A}}_\varepsilon^{\,1/2})$ a similar result was obtained by Meshkova [26], [27]:
where $K(\tau,\varepsilon)$ is an appropriate corrector. In the manuscript [28] an approximation for the operator $\widehat{\mathcal{A}}_\varepsilon^{-1/2} \sin(\tau \widehat{\mathcal{A}}_\varepsilon^{\,1/2})$ in the $(H^3 \to L_2)$-norm was found with corrector taken into account, the error term being of order $O(\varepsilon^2)$. Results with correctors were obtained in [26]–[28] due to the presence of the ‘smoothing’ factor $\widehat{\mathcal{A}}_\varepsilon^{-1/2}$ in the operator to be approximated. No analogues of such results were previously known for the operators $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ and $\cos(\tau \widehat{\mathcal{A}}_\varepsilon^{\,1/2})$.
Let us explain the method by using the example of the proof of estimate (0.7). Set $\mathcal{H}_0:=-\Delta$. Clearly, estimate (0.7) is equivalent to the inequality
Next, applying the unitary Gelfand transform, we decompose the operator $\widehat{\mathcal{A}}$ into the direct integral of the operators $\widehat{\mathcal{A}}(\mathbf k)$ acting in $L_2(\Omega;\mathbb{C}^n)$ (where $\Omega$ is a cell of the lattice $\Gamma$) and given by the expression $b({\mathbf D}+\mathbf k)^* g({\mathbf x}) b({\mathbf D}+\mathbf k)$ with periodic boundary conditions. The operators $\widehat{\mathcal{A}}(\mathbf k)$ have discrete spectra. The operator family $\widehat{\mathcal{A}}(\mathbf k)$ is studied using methods of analytic perturbation theory (with respect to the one-dimensional parameter $t=|\mathbf k|$). It is possible to obtain an analogue of inequality (0.12) for the operators $\widehat{\mathcal{A}}(\mathbf k)$ with a constant independent of $\mathbf k$. This yields estimate (0.12).
The papers [29] and [30] were devoted to the further study of the operator exponential. In [29] it was shown that estimate (0.7) is sharp with respect to the type of the operator norm: the conditions on the operator were specified under which the estimate
is not true if $s<3$. In [30] it was shown that estimate (0.7) is also sharp with respect to the dependence on $\tau$ (for large $|\tau|$): the factor $(1+|\tau|)$ on the right-hand side of the estimate cannot be replaced by $(1+|\tau|)^{\alpha}$ for $\alpha<1$. On the other hand, in [29] additional conditions on the operator were found under which the result was improved with respect to the type of the norm: $H^3$ can be replaced by $H^2$. In [30] it was shown that under the same conditions the result admits improvement in another sense: the factor $(1+|\tau|)$ can be replaced by $(1+|\tau|)^{1/2}$. As a result, under certain additional assumptions (which are automatically fulfilled for the acoustics operator), the following estimate was proved:
$$
\begin{equation*}
\| e^{- i \tau \widehat{\mathcal{A}}_\varepsilon}- e^{-i \tau \widehat{\mathcal{A}}^{\,0}}\|_{H^2(\mathbb{R}^d) \to L_2(\mathbb{R}^d)}\leqslant C (1+|\tau|)^{1/2}\varepsilon.
\end{equation*}
\notag
$$
Hyperbolic problems were studied in [31] and [32]. It was shown there that estimates (0.8)–(0.10) are sharp both with respect to the type of the operator norm, and with respect to the dependence on $\tau$. However, under certain additional assumptions these results can be improved in both senses.
Non-stationary problems were also investigated for the more general class of operators $\widehat{\mathcal B}_\varepsilon$ (with lower-order terms): the exponential $e^{-i \tau \widehat{\mathcal{B}}_\varepsilon}$ was studied in [33], and hyperbolic problems were considered in [28] and [34]. Moreover, in [34] another approach to hyperbolic problems was proposed, based on a modification of the Trotter–Kato theorem.
0.4. The development of operator estimates in homogenization theory
Let us briefly discuss other directions of study of operator error estimates in homogenization problems.
Using the operator-theoretic approach, operator estimates were obtained for homogenization of the stationary and non-stationary Maxwell systems in $\mathbb{R}^3$ (see [35]–[37]). Recently, this approach has been adapted to the study of homogenization of non-local convolution-type operators [38]. (Such operators arise in models of mathematical biology and population dynamics and have actively been studied recently; see, for instance, [39]–[41].)
Operator estimates were studied for elliptic operators in $\mathbb{R}^d$ of higher even order. The operator-theoretic approach was applied to matrix higher-order operators in [42]–[46]. This approach was also used to study homogenization of the parabolic higher-order equations in [47]. Homogenization of Schrödinger-type equations and hyperbolic equations with higher-order operators was studied in [48] and [49].
The shift method was applied to homogenization of higher-order operators by Pastukhova; see [50]–[54] and the references there.
Operator estimates were studied not only for the homogenization problem for the elliptic operator ${\mathcal A}_\varepsilon$ in $\mathbb{R}^d$, but also for boundary-value problems in a bounded domain. In [22], for second-order operators under the Dirichlet or Neumann conditions, operator error estimates of order $O(\varepsilon^{1/2})$ in the $(L_2 \to L_2)$- and $(L_2 \to H^1)$- norms were obtained; estimates deteriorate due to the influence of the boundary. Close results were obtained by Griso [55], [56] for a scalar elliptic operator in a bounded domain under the Dirichlet or Neumann conditions with the help of the unfolding method; in [56], a sharp order error estimate (of order $O(\varepsilon)$) for approximation of the resolvent in the operator norm on $L_2$ was proved for the first time. Similar results for elliptic systems were independently obtained in [57] and [58]–[60]. Further results were found in [61]–[65]; see also the monograph [66] by Shen and the references there. Operator estimates for homogenization of the initial boundary-value problems for parabolic equations were studied in [67]–[69]. For a stationary Maxwell system in a bounded domain under the perfect conductivity boundary conditions, such estimates were found in [70], [71], and for higher-order operators in a bounded domain in [72]–[75].
In recent years operator error estimates in various homogenization problems for differential operators have been attracting the attention of an increasing number of researchers; many meaningful results have been obtained. Such estimates have been studied for the Stokes system [76], for operators with locally periodic and multiscale coefficients [77]–[87], in problems with high contrast [88], [89], in problems with rapidly oscillating boundary or frequently changing type of boundary conditions [90]–[93]. Many papers are devoted to operator estimates in problems with perforation; see [94] and [95], where the spectral approach was used, as well as [96]–[104]. Here we do not touch on results on operator estimates for nonlinear equations, or equations with almost periodic or random coefficients, and do not pretend on completeness of the survey.
0.5. Main results
In the present paper we give a survey of the known results on operator estimates for homogenization of Schrödinger-type equations and also obtain new results on the behaviour of the operator exponential $e^{-i\tau\widehat{\mathcal{A}}_\varepsilon}$ for small $\varepsilon$. We are interested in whether it is possible, for fixed $\tau$, to find approximations for the exponential $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ in the $(H^s \to L_2)$-norm (for suitable $s$) with an error of $O(\varepsilon^2)$ and in the $(H^s \to H^1)$-norm with an an error of $O(\varepsilon)$. It is impossible to construct such approximations for the exponential $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ itself. Instead, we find such approximations for the ‘corrected’ exponential, which is the composition of the operators $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ and $I+\varepsilon \Lambda^\varepsilon b({\mathbf D}) \Pi_\varepsilon$. Here $\Lambda({\mathbf x})$ is the periodic solution of the cell problem (see (6.8)), and $\Pi_\varepsilon$ is an auxiliary smoothing operator.
Here ${\mathcal K}(\varepsilon)$ and ${\mathcal K}_1(\varepsilon)$ are appropriate correctors; they contain the rapidly oscillating coefficient $\Lambda^\varepsilon$, and therefore depend on $\varepsilon$. The effective operator and the correctors are described in terms of the spectral characteristics of the operator $\widehat{\mathcal{A}}$ at the bottom of the spectrum. It is impossible to approximate the exponential $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ itself in the same terms with the required accuracy, because the ‘problematic’ term $e^{-i\tau\widehat{\mathcal{A}}_\varepsilon} \varepsilon\Lambda^\varepsilon b({\mathbf D})\Pi_\varepsilon$ cannot be approximated in threshold terms; see the discussion in § 14.6.
On the one hand, we confirm that estimates (0.13), (0.14) are sharp: a condition on the operator is given under which these estimates cannot be improved either with respect to the type of the operator norm or with respect to the dependence on $\tau$. This condition is formulated in spectral terms.
We consider the operator family $\widehat{\mathcal A}(\mathbf k)$ and put $\mathbf k=t \boldsymbol{\theta}$, $t=|\mathbf k|$, $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. This family is analytic with respect to the parameter $t$. For $t=0$ the point $\lambda_0=0$ is an $n$-multiple eigenvalue of the ‘unperturbed’ operator $\widehat{\mathcal A}(0)$. Then for small $t$ there exist real-analytic branches of eigenvalues $\lambda_l(t,\boldsymbol{\theta})$ ($l=1,\dots,n$) of the operator $\widehat{\mathcal A}(\mathbf k)$. For small $t$ the following convergent power series expansions are valid:
where $\gamma_l(\boldsymbol{\theta}) >0$ and $\mu_l(\boldsymbol{\theta}),\nu_l(\boldsymbol{\theta}) \in \mathbb{R}$. The condition under which estimates (0.13) and (0.14) cannot be improved is that $\mu_l(\boldsymbol{\theta}_0)\ne 0$ for some $l$ and some $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$.
On the other hand, under certain additional assumptions we improve the results and obtain the estimates
For $n=1$ a sufficient condition that ensures estimates (0.15) and (0.16) is that $\mu_1(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. In particular, this condition holds for the operator $\widehat{\mathcal{A}}_\varepsilon={\mathbf D}^* g^\varepsilon({\mathbf x}){\mathbf D}$ if $g({\mathbf x})$ is a symmetric matrix with real entries. For $n\geqslant 2$, to ensure (0.15) and (0.16), in addition to the condition that all coefficients $\mu_l(\boldsymbol{\theta})$ are equal to zero, we impose one more condition in terms of the coefficients $\gamma_l(\boldsymbol{\theta})$. The simplest version of this condition is that different branches $\gamma_l(\boldsymbol{\theta})$ do not intersect each other.
Next, we show that estimates (0.15) and (0.16) are also sharp: if all coefficients $\mu_l(\boldsymbol{\theta})$ are equal to zero, but $\nu_j(\boldsymbol{\theta}_0) \ne 0$ (for some $j$ and some $\boldsymbol{\theta}_0$), then estimates (0.15) and (0.16) cannot be improved either with respect to the norm type, or with respect to the dependence on $\tau$.
Using interpolation, we also obtain estimates in the $(H^s \to L_2)$- or $(H^s \to H^1)$- norm. For instance, the $(H^s \to L_2)$-norm of the operator from (0.13) satisfies estimate of order $O((1+|\tau|)^{s/3}\varepsilon^{s/3})$ for $3\leqslant s \leqslant 6$. In the case of improvement, the $(H^s \to L_2)$-norm of this operator is $O((1+|\tau|)^{s/4}\varepsilon^{s/2})$ for $2\leqslant s \leqslant 4$.
Clearly, the results obtained yield qualified error estimates for small $\varepsilon$ and large $\tau$: in the general case, it is possible to consider $\tau =O(\varepsilon^{-\alpha})$ for $0<\alpha<1$, while in the case of improvement it is possible to consider $\tau=O(\varepsilon^{-\alpha})$ for $0<\alpha<2$.
In the case of a more general operator (0.3) we obtain analogues of the results described above for the ‘sandwiched’ operator exponential $f^\varepsilon e^{-i\tau{\mathcal{A}}_\varepsilon}(f^\varepsilon)^{-1}$.
The results formulated in operator terms are applied to homogenization of the solutions of the Cauchy problem for the Schrödinger-type equations with initial data from a special class. In particular, we consider the non-stationary Schrödinger equation and the two-dimensional Pauli equation with singular rapidly oscillating potentials. Note that it is not rare when one cannot find asymptotics (with respect to some parameter) of the solutions of the Cauchy problem for non-stationary Schrödinger-type equations or hyperbolic equations for all initial data, but only for initial data in some special class; see, for example, [105]–[107].
Similar results were obtained by Dorodnyi and this author for homogenization of hyperbolic equations; a brief communication on these results was published in [108]; a detailed paper is under preparation.
0.6. The method
The results we discuss are obtained by a further development of the operator-theoretic approach. We follow the plan outlined in § 0.3 above. Our considerations are based on an abstract operator-theoretic scheme. A family of operators $A(t)=X(t)^*X(t)$, $t \in \mathbb{R}$, acting on some Hilbert space ${\mathfrak H}$ is studied. Here $X(t)=X_0+tX_1$. (The family $A(t)$ models the operator family ${\mathcal A}(\mathbf k)={\mathcal A}(t \boldsymbol{\theta})$, but in the abstract setting the parameter $\boldsymbol{\theta}$ is absent.) It is assumed that the point $\lambda_0 =0$ is an isolated eigenvalue of multiplicity $n$ for the operator $A(0)$. Then for $|t| \leqslant t_0$ the perturbed operator $A(t)$ has exactly $n$ eigenvalues on an interval $[0,\delta]$ ($\delta$ and $t_0$ are controlled explicitly). These eigenvalues and the corresponding eigenvectors are real-analytic functions of $t$. The coefficients of the corresponding power series expansions are called the threshold characteristics of the operator $A(t)$. We distinguish a finite-rank operator $S$ (the so-called spectral germ of the family $A(t)$) acting on the subspace ${\mathfrak N}=\operatorname{Ker} A(0)$. The spectral germ carries information about the threshold characteristics of principal order.
In terms of the spectral germ we find the principal term of approximation of the operator $e^{-i\varepsilon^{-2}\tau A(t)}$. To find more accurate approximations with correctors, we need to take into account the threshold characteristics of the next order. Applying these abstract results leads to the required estimates for differential operators.
The present paper relies on the abstract material prepared in [109].
0.7. The plan of the paper
The paper consists of three chapters. In Chapter 1 (§§ 1–4), we briefly present the necessary abstract operator-theoretic material.
In Chapter 2 (§§ 5–13), periodic differential operators of the form (0.1), (0.2) are studied. In § 5 the class of operators under consideration is introduced and the direct integral decomposition is described; the corresponding operator family ${\mathcal A}(\mathbf k)$ is included in the framework of the abstract scheme. The effective characteristics of the operator $\widehat{\mathcal A}$ are described in § 6. In § 7, using abstract theorems, we deduce approximations for the operator $e^{-i \varepsilon^{-2} \tau \widehat{\mathcal A}(\mathbf k)}$, and in § 8 we confirm the sharpness of these results. In § 9 the effective characteristics of the operator (0.1) are described. The required approximations for the sandwiched exponential of ${\mathcal A}(\mathbf k)$ are found in § 10, and the sharpness of these results is discussed in § 11. Section 12 is devoted to approximations for the exponential $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal A}}$ of the operator (0.2), and in § 13 we find the required approximations for the sandwiched exponential of the operator (0.1). These results are deduced from the results of §§ 7–11 by means of the direct integral decompositions.
Chapter 3 (§§ 14–19) is devoted to homogenization problems. In §§ 14 and 15, with the help of the scaling transformation, we deduce the main results of the paper (approximations for the exponential $e^{-i\tau\widehat{{\mathcal A}}_\varepsilon}$ and for the sandwiched exponential $e^{-i\tau {\mathcal A}_\varepsilon}$) from the results of Chapter 2. In § 16 the results obtained are applied to the investigation of the solutions of the Cauchy problem for Schrödinger-type equations. Sections 17–19 are devoted to applications of the general results to particular equations of mathematical physics.
0.8. The notation
Let ${\mathfrak H}$ and ${\mathfrak H}_*$ be complex separable Hilbert spaces. The symbols $(\,\cdot\,{,}\,\cdot\,)_{\mathfrak H}$ and $\|\,{\cdot}\,\|_{\mathfrak H}$ denote the inner product and norm in ${\mathfrak H}$, respectively; the symbol $\|\,{\cdot}\,\|_{{\mathfrak H} \to {\mathfrak H}_*}$ denotes the norm of a bounded operator from ${\mathfrak H}$ to ${\mathfrak H}_*$. Sometimes we omit indices. By $I=I_{\mathfrak H}$ we denote the identity operator in ${\mathfrak H}$. If $A\colon {\mathfrak H} \to {\mathfrak H}_*$ is a linear operator, then $\operatorname{Dom} A$ and $\operatorname{Ker} A$ denote its domain and its kernel, respectively. If ${\mathfrak N}$ is a subspace of ${\mathfrak H}$, then ${\mathfrak N}^\perp$ is its orthogonal complement. If $P$ is the orthogonal projection of ${\mathfrak H}$ onto ${\mathfrak N}$, then $P^\perp$ is the orthogonal projection onto ${\mathfrak N}^\perp$.
The symbols $\langle\,{\cdot}\,{,}\,{\cdot}\,\rangle$ and $|\,{\cdot}\,|$ denote the inner product and norm in $\mathbb{C}^n$; ${\mathbf 1}_n$ is the identity $n \times n$ matrix. If $a$ is an $ m\times n $ matrix, then the symbol $|a|$ denotes the norm of the matrix $a$ viewed as a linear operator from $\mathbb{C}^n$ to $\mathbb{C}^m$.
Next, we set ${\mathbf x}=(x_1,\dots,x_d) \in \mathbb{R}^d$, $i D_j=\partial_j=\partial/ \partial x_j$, $j=1,\dots,d$, and ${\mathbf D}=-i\nabla=(D_1,\dots,D_d)$.
The classes $L_p$ (where $1 \leqslant p \leqslant \infty$) of $\mathbb{C}^n$-valued functions in a domain ${\mathcal O} \subset \mathbb{R}^d$ are denoted by $L_p({\mathcal O};\mathbb{C}^n)$. The Sobolev classes of order $s$ (where $s \geqslant 0$) of $\mathbb{C}^n$-valued functions in a domain ${\mathcal O}$ are denoted by $H^s({\mathcal O};\mathbb{C}^n)$. If $n=1$, then we write simply $L_p({\mathcal O})$ and $H^s({\mathcal O})$, but sometimes we use this simple notation also for classes of vector-valued or matrix-valued functions.
Different constants in estimates are denoted by $C$, $\mathcal C$, $\mathfrak C$, and $c$ (probably, with indices and marks).
Chapter 1. Abstract operator-theoretic scheme
This chapter contains abstract material borrowed from [7], [8], [15], [29], [30], and [109].
1. Quadratic operator pencils
1.1. The operators $X(t)$ and $A(t)$
Let $\mathfrak{H}$ and $\mathfrak{H}_{*}$ be complex separable Hilbert spaces. Suppose that $X_{0}\colon \mathfrak{H} \to \mathfrak{H}_{*}$ is a densely defined and closed operator and $X_{1}\colon \mathfrak{H} \to \mathfrak{H}_{*}$ is a bounded operator. Then the operator $X(t)=X_0+t X_1$, $t \in \mathbb{R}$, is closed on $\operatorname{Dom} X_0$. Consider the family of selfadjoint operators $A(t)=X(t)^* X(t)$ in $\mathfrak{H}$. The operator $A(t)$ is generated by the closed quadratic form $\|X(t)u\|^{2}_{\mathfrak{H}_*}$, $u \in \operatorname{Dom} X_0$. Denote $A_0:=A(0)$, $\mathfrak{N}:=\operatorname{Ker} A_0=\operatorname{Ker} X_0$, and $\mathfrak{N}_{*}:=\operatorname{Ker} X^*_0$.
It is assumed that the point $\lambda_0=0$ is an isolated point of the spectrum of $A_0$, $0 < n:=\dim \mathfrak{N} < \infty$, and $n \leqslant n_*:=\dim \mathfrak{N}_* \leqslant \infty.$
Let $d^0$ be the distance of the point $\lambda_0=0$ to the rest of the spectrum of $A_0$. By $P$ and $P_*$ we denote the orthogonal projections of $\mathfrak{H}$ onto $\mathfrak{N}$ and of $\mathfrak{H}_*$ onto $\mathfrak{N}_*$, respectively. Let $F(t;[a, b])$ be the spectral projection of the operator $A(t)$ for the interval $[a,b]$. We put $\mathfrak{F}(t;[a,b]):=F(t;[a, b])\mathfrak{H}$. Fix a number $\delta > 0$ such that $8 \delta < d^0$. We choose a number $t_0 > 0$ so that
As shown in [7], Chap. 1, Proposition 1.2, for $|t| \leqslant t_0$ we have $F(t;[0,\delta])=F(t;[0,3\delta])$ and $\operatorname{rank} F(t;[0,\delta])=n$. We will write $F(t)$ instead of $F(t;[0,\delta])$ and $\mathfrak{F} (t)$ instead of $\mathfrak{F}(t;[0,\delta])$.
1.2. Auxiliary operators
In accordance with [7], Chap. 1, § 1, and [8], § 1, we introduce the operators appearing in our considerations of perturbation theory.
Denote $\mathcal{D}:=\operatorname{Dom} X_0 \cap \mathfrak{N}^{\perp}$. Let $\omega \in \mathfrak{N}$. Consider the equation
There exists a unique solution $\phi=\phi(\omega)$. We introduce the operator $Z\colon \mathfrak{H} \to \mathfrak{H}$ by the relation $Zu=\phi(Pu)$, $u \in \mathfrak{H}$. Note that $PZ=0$, hence $Z^* P=0$. We have
Next we define the operator $R \colon \mathfrak{N} \to \mathfrak{N}_*$ by the formula $R:=X_0 Z+X_1$. Another representation for $R$ is given by $R= P_*X_1\big|_{\mathfrak{N}}$.
According to [7], Chap. 1, the operator $S:=R^* R\colon \mathfrak{N} \to \mathfrak{N}$ is called the spectral germ of the family $A(t)$ at $t=0$. The germ can be represented as $S=P X^*_1 P_* X_1 \big|_{\mathfrak{N}}$. The spectral germ is called non-degenerate if $\operatorname{Ker} S=\{0\}$. Note that $\| R \| \leqslant \| X_1 \|$ and $\| S \| \leqslant \| X_1 \|^2$.
We also need the operators $Z_2$ and $R_2$ (see [15], § 1). Let $\omega \in \mathfrak{N}$, and let $\psi= \psi(\omega) \in \mathcal{D}$ be a (weak) solution of the equation $X^*_0(X_0\psi+X_1Z\omega)=-P^\perp X_1^*R\omega$. Obviously, the solvability condition is satisfied. We define the operator $Z_2\colon \mathfrak{H} \to \mathfrak{H}$ by $Z_2 u=\psi(P u)$, $u \in \mathfrak{H}$. Finally, we introduce the operator $R_2\colon \mathfrak{N} \to \mathfrak{H}_*$ by the formula $R_2:=X_0 Z_2+X_1 Z$.
1.3. Analytic branches of eigenvalues and eigenvectors of the operator $A(t)$
According to the general analytic perturbation theory (see [110]), for $|t| \leqslant t_0$ there exist real-analytic functions $\lambda_l (t)$ (branches of eigenvalues) and real-analytic $\mathfrak{H}$-valued functions $\varphi_l (t)$ (branches of eigenvectors) such that
and the set $\varphi_l (t)$, $l=1,\dots,n$, forms an orthonormal basis in $\mathfrak{F}(t)$. For sufficiently small $t_*$ (where $0 < t_* \leqslant t_0$) and $|t| \leqslant t_*$ we have the following convergent power series expansions:
The elements $\omega_l= \varphi_l (0)$, $l=1,\dots,n$, form an orthonormal basis in $\mathfrak{N}$. Substituting expansions (1.4) and (1.5) into (1.3) and comparing the coefficients of $t$ and $t^2$, we arrive at the following relations:
$$
\begin{equation}
S \omega_l=\gamma_l \omega_l, \qquad l=1,\dots,n
\end{equation}
\tag{1.7}
$$
(cf. [7], Chap. 1, § 1, and [8], § 1). Thus, the numbers $\gamma_l$ and the elements $\omega_l$ defined by (1.4) and (1.5) are the eigenvalues and eigenvectors of the germ $S$. We have
The spectral projection $F(t)$ and the operator $A(t)F(t)$ are real-analytic operator-valued functions for $|t| \leqslant t_0$. Combining the representations
converging for $|t| \leqslant t_*$. However, we do not need expansions, but only approximations (with one or several first terms) with error estimates on the controlled interval $|t| \leqslant t_0$.
The following statement was obtained in [7] (see [7], Chap. 1, Theorems 4.1 and 4.3). Below we denote by $\beta_j$ absolute constants assuming that $\beta_j \geqslant 1$.
Proposition 1.1 ([7]). Under the assumptions of § 1.1 we have
where $K_0$ takes $\mathfrak{N}$ to $\mathfrak{N}^{\perp}$ and $\mathfrak{N}^{\perp}$ to $\mathfrak{N}$, and $N=N_0+N_*$ takes $\mathfrak{N}$ to itself and takes $\mathfrak{N}^{\perp}$ to $\{0\}$. In terms of the coefficients of power series expansions the operators $F_1$, $K_0$, $N_0$, and $N_*$ are given by
Remark 1.3. In the basis $\{\omega_l\}_{l=1}^n$ the operators $N$, $N_0$, and $N_*$ (as restricted to $\mathfrak{N}$) are given by matrices of size $n \times n$. The operator $N_0$ is diagonal:
From (1.16) it follows that $\lambda_l (t) \geqslant c_* t^2$, $l=1,\dots,n$, for $|t| \leqslant t_0$. By (1.4) this implies that $\gamma_l \geqslant c_* > 0$, $l=1,\dots,n$. Thus, the spectral germ is non-degenerate:
$$
\begin{equation}
S \geqslant c_* I_{\mathfrak{N}}.
\end{equation}
\tag{1.17}
$$
1.6. Dividing the eigenvalues of the operator $A(t)$ into clusters
The material of this subsection is borrowed from [29], § 2. It is meaningful for $n \geqslant 2$.
Suppose that Condition 1.4 is satisfied. Now it will be convenient to change the notation, tracing the multiplicities of eigenvalues of the germ $S$. Let $p$ be the number of different eigenvalues of the germ. We enumerate these eigenvalues in the increasing order and denote them by $\gamma_j^\circ$, $j=1,\dots,p$. Their multiplicities are denoted by $k_1,\dots,k_p$ (obviously, $k_1+\cdots+k_p=n$). The eigenspaces are denoted by $\mathfrak{N}_j:=\operatorname{Ker}(S-\gamma^{\circ}_j I_\mathfrak{N})$, $j=1,\dots,p$. Then $\mathfrak{N}=\bigoplus_{j=1}^p\mathfrak{N}_j$. Let $P_j$ be the orthogonal projection of $\mathfrak{H}$ onto $\mathfrak{N}_j$. Then $P= \sum_{j=1}^{p}P_j$ and $P_j P_l=0$ for $j \ne l$. Correspondingly, we change the notation for eigenvectors of the germ (which are ‘embryos’ in (1.5)) by dividing them into $p$ groups, so that $\omega^{(j)}_1,\dots,\omega^{(j)}_{k_j}$ correspond to the eigenvalue $\gamma^{\circ}_j$ and form an orthonormal basis in $\mathfrak{N}_j$.
Remark 1.5. Recall that $N=N_0+N_*$. According to Remark 1.3, $P_j N_* P_j=0,$ $j=1,\dots,p,$ and $P_l N_0 P_j =0$ for $l \ne j$. This yields invariant representations for the operators $N_0$ and $N_*$:
$$
\begin{equation}
N_0=\sum_{j=1}^{p} P_j N P_j \quad\text{and} \quad N_*=\sum_{\substack{1 \leqslant j, l \leqslant p: \\ j \ne l}} P_j N P_l.
\end{equation}
\tag{1.18}
$$
For each pair of indices $(j,l)$, $1 \leqslant j,l \leqslant p$, $j \ne l$, we denote
Clearly, there exists a number $i_0=i_0(j,l)$, where $j \leqslant i_0 \leqslant l-1$ for $j < l$ and $l \leqslant i_0 \leqslant j-1$ for $l < j$, such that $\gamma^{\circ}_{i_0+1}-\gamma^{\circ}_{i_0} \geqslant c^{\circ}_{jl}$. This means that on the interval between $\gamma^{\circ}_j$ and $\gamma^{\circ}_l$ there is a gap in the spectrum of $S$ of length at least $c^{\circ}_{jl}$. The choice of $i_0$ can be ambiguous; in this case we agree (for certainty) to take the smallest possible $i_0$.
We choose a number $t^{00}_{jl} \leqslant t_0$ satisfying the inequality
Denote $\Delta_{jl}^{(1)}:= [\gamma^{\circ}_1-c^{\circ}_{jl}/4,\gamma^{\circ}_{i_0}+c^{\circ}_{jl}/4]$ and $\Delta_{jl}^{(2)}:=[\gamma^{\circ}_{i_0+1}- c^{\circ}_{jl}/4,\gamma^{\circ}_p+c^{\circ}_{jl}/4]$. The distance between the intervals $\Delta_{jl}^{(1)}$ and $\Delta_{jl}^{(2)}$ is not less than $c^{\circ}_{jl}/2$. In [29], § 2, it was shown that for $|t| \leqslant t^{00}_{jl}$ the operator ${A}(t)$ has exactly $k_1+\cdots+k_{i_0}$ eigenvalues (with multiplicities taken into account) on the interval $t^2 \Delta_{jl}^{(1)}$ and exactly $k_{i_0+1}+\cdots+k_p$ eigenvalues on the interval $t^2 \Delta_{jl}^{(2)}$.
1.7. The coefficients $\nu_l$
For definiteness suppose that the enumeration in (1.4) and (1.5) is such that $\gamma_1 \leqslant \cdots \leqslant \gamma_n$. The coefficients $\nu_l$ and the vectors $\omega_l$, $l=1,\dots,n$, in the expansions (1.4) and (1.5) are eigenvalues and eigenvectors of some problem; see [30], § 1.8. We need to describe this problem in the case where $\mu_l=0$ and $l=1,\dots,n$, that is $N_0=0$; see also [32], Proposition 1.7.
Using the notation of § 1.6, we introduce operators $\mathcal{N}^{(q)}$, $q=1,\dots,p$ : the operator $\mathcal{N}^{(q)}$ acts in $\mathfrak{N}_q$ and is given by
$$
\begin{equation*}
\mathcal{N}^{(q)}:=P_q\biggl(N_1^0-\frac{1}{2} Z^* Z SP- \frac{1}{2}SP Z^* Z\biggr)\biggl|_{\mathfrak{N}_q}+ \sum_{j=1,\dots,p: j\ne q}(\gamma_q^\circ-\gamma_j^\circ)^{-1} P_q N P_j N\big|_{\mathfrak{N}_q}.
\end{equation*}
\notag
$$
Denote $i(q)=k_1+\cdots+k_{q-1}+1$. Let $\nu_l$, $l=1,\dots,n,$ be the coefficients of $t^4$ in the expansions (1.4). Then
2. Approximation for the operator $e^{-i\varepsilon^{-2} \tau A(t)}$
2.1. Approximation in the operator norm on ${\mathfrak H}$
Now we introduce a small parameter $\varepsilon > 0$ and describe the behaviour of the operator $e^{-i\varepsilon^{-2} \tau A(t)}$ for small $\varepsilon$. It is convenient to multiply this operator by the ‘smoothing factor’ $\varepsilon^s (t^2+\varepsilon^2)^{-s/2}P$, where $s > 0$. (This term is explained by the fact that in applications to differential operators such multiplication turns into smoothing.)
In [25], Theorem 2.6, the following result was obtained.
Theorem 2.1 ([25]). For $\varepsilon > 0$, $\tau \in \mathbb{R}$, and $|t| \leqslant t_0$ we have
Under some additional assumptions this result can be improved; see [30], Theorems 2.5 and 2.6. Recall that the operator $N$ is defined by (1.13), and $N_0$ is defined by (1.18).
Theorem 2.2 ([30]). Suppose that $N=0$. Then for $\varepsilon > 0$, $\tau \in \mathbb{R}$, and $|t| \leqslant t_0$ we have
Remark 2.10. In [25], [30], and [109], explicit expressions for the constants in estimates from Theorems 2.1–2.9 were found. The following is essential. The constants $C_1$, $C_{2}$, $C_6$, $C_{9}$, $C_{10}$, $C_{11}$, $C_{12}$, $C_{15}$, $C_{16}$, and $C_{17}$ in Theorems 2.1, 2.2, 2.4, 2.5, 2.7, and 2.8 are estimated by polynomials with (absolute) positive coefficients in the variables $\delta^{-1/2}$ and $\|X_1\|$. The constants $C_{7}$, $C_{8}$, $C_{13}$, $C_{14}$, $C_{18}$, and $C_{19}$ in Theorems 2.3, 2.6, and 2.9 are estimated by polynomials with positive coefficients in the same variables, and also in $(c^{\circ})^{-1}$ and $n$. Here $c^{\circ}$ is the constant defined by (1.19) and (2.1).
2.4. Confirming sharpness with respect to the smoothing factor
The following theorem confirms that in the general case Theorems 2.1, 2.4, and 2.7 are sharp with respect to the smoothing factor.
Theorem 2.11 ([29], [109]). Suppose that $N_0 \ne 0$.
$1^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 3$. Then there does not exist a constant $C(\tau)$ such that the inequality
holds for sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [29], Theorem 4.4, statement $2^\circ$ in [109], Theorem 4.3, and statement $3^\circ$ in [109], Theorem 4.6.
Further, it turns out that Theorems 2.2, 2.3, 2.5, 2.6, 2.8, and 2.9 (on improvements of general results under additional assumptions), in their turn, are sharp. Recall that the operator $\mathcal{N}^{(q)}$ was defined in § 1.7.
Theorem 2.12 ([30], [109]). Suppose that $N_0=0$ and $\mathcal{N}^{(q)} \ne 0$ for some $q\in \{1,\dots,p\}$.
$1^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 2$. Then there does not exist a constant $C(\tau)$ such that inequality (2.3) holds for sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 4$. Then there does not exist a constant $C(\tau)$ such that inequality (2.4) holds for sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 3$. Then there does not exist a constant $C(\tau)$ such that inequality (2.5) holds for sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was obtained in [30], Theorem 2.9, statement $2^\circ$ in [109], Theorem 4.4, and statement $3^\circ$ in [109], Theorem 4.7.
2.5. Sharpness of results with respect to time
The following theorem confirms that in the general case Theorems 2.1, 2.4, and 2.7 are sharp with respect to the dependence of estimates on $\tau$ (for large $|\tau|$).
Theorem 2.13 ([30], [109]). Suppose that $N_0 \ne 0$.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau|=0$ and estimate (2.3) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ \tau^2 =0$ and estimate (2.4) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau| =0$ and estimate (2.5) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [30], Theorem 2.10, statement $2^\circ$ in [109], Theorem 4.10, and statement $3^\circ$ in [109], Theorem 4.13.
Further, Theorems 2.2, 2.3, 2.5, 2.6, 2.8, and 2.9 (on improvements of general results under additional assumptions), in their turn, are sharp.
Theorem 2.14 ([30], [109]). Suppose that $N_0=0$ and $\mathcal{N}^{(q)} \ne 0$ for some $q\in \{1,\dots,p\}$.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau|^{1/2} =0$ and estimate (2.3) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau| =0$ and estimate (2.4) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau|^{1/2} =0$ and estimate (2.5) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [30], Theorem 2.11, statement $2^\circ$ in [109], Theorem 4.11, and statement $3^\circ$ in [109], Theorem 4.14.
3. Operator of the form $A(t)=M^*\widehat{A}(t)M$
3.1. The operator family of the form $A(t)=M^*\widehat{A}(t)M$
Along with the space $\mathfrak{H}$, we consider yet another separable Hilbert space $\widehat{\mathfrak{H}}$. Let $\widehat{X} (t)=\widehat{X}_0+t \widehat{X}_1 \colon \widehat{\mathfrak{H}} \to \mathfrak{H}_*$ be a family of operators of the same form as $X(t)$. Suppose that $\widehat{X}(t)$ satisfies the assumptions of § 1.1. Let $M \colon \mathfrak{H} \to \widehat{\mathfrak{H}}$ be an isomorphism. Assume that $M \operatorname{Dom} X_0=\operatorname{Dom} \widehat{X}_0$ and $X(t)=\widehat{X} (t) M$, and then also $X_0=\widehat{X}_0 M$ and $X_1=\widehat{X}_1 M$. In $\widehat{\mathfrak{H}}$ we introduce the family of selfadjoint operators $\widehat{A}(t)=\widehat{X}(t)^*\widehat{X}(t)$. Then, obviously,
In what follows all objects corresponding to the family $\widehat{A}(t)$ are marked by hats ‘$\, \widehat{\phantom{\_}} \,$’. Note that $\widehat{\mathfrak{N}}=M \mathfrak{N}$, $\widehat{n}=n$, $\widehat{\mathfrak{N}}_*= \mathfrak{N}_*$, $\widehat{n}_*=n_*$, and $\widehat{P}_*=P_*$.
In the space $\widehat{\mathfrak{H}}$ we consider the positive definite operator
Let $Q_{\widehat{\mathfrak{N}}}$ be the block of the operator $Q$ in $\widehat{\mathfrak{N}}$, that is, $Q_{\widehat{\mathfrak{N}}}=\widehat{P} Q\big|_{\widehat{\mathfrak{N}}} \colon \widehat{\mathfrak{N}} \to \widehat{\mathfrak{N}}$. Obviously, $Q_{\widehat{\mathfrak{N}}}$ is an isomorphism in $\widehat{\mathfrak{N}}$.
As shown in [12], Proposition 1.2, the orthogonal projection $P$ of $\mathfrak{H}$ onto $\mathfrak{N}$ and the orthogonal projection $\widehat{P}$ of $\widehat{\mathfrak{H}}$ onto $\widehat{\mathfrak{N}}$ satisfy the following relation:
Let $\widehat{S}\colon\widehat{\mathfrak{N}}\to\widehat{\mathfrak{N}}$ be the spectral germ of the family $\widehat{A}(t)$ at $t=0$, and let $S$ be the germ of the family $A(t)$. In [7], Chap. 1, § 1.5, it was proved that
Assume that $A(t)$ satisfies Condition 1.4. Then the germ $S$ (as well as $\widehat{S}$) is non-degenerate.
3.2. The operators $\widehat{Z}_Q$ and $\widehat{N}_Q$
For the operator family $\widehat{A}(t)$ we introduce the operator $\widehat{Z}_Q$ acting in $\widehat{\mathfrak{H}}$ and taking an element $\widehat{u} \in \widehat{\mathfrak{H}}$ to the solution $\widehat{\phi}_Q$ of the problem
where $\widehat{\omega}=\widehat{P} \widehat{u}$. As shown in [8], § 6, the operator $Z$ for the family $A(t)$ and the operator $\widehat{Z}_Q$ introduced above satisfy the following relation:
$$
\begin{equation}
\widehat{Z}_Q =M Z M^{-1} \widehat{P}.
\end{equation}
\tag{3.4}
$$
Then $\widehat{N}_Q=\widehat{N}_{0,Q}+\widehat{N}_{*,Q}$.
The following lemma was proved in [29], Lemma 5.1.
Lemma 3.1 ([29]). Suppose that the assumptions of § 3.1 are satisfied. Let $N$ and $N_0$ be the operators defined by (1.13) and (1.18). Suppose that the operators $\widehat{N}_Q$ and $\widehat{N}_{0,Q}$ are defined by (3.6) and (3.10). Then the condition $N=0$ is equivalent to the relation $\widehat{N}_Q=0$. The condition $N_0=0$ is equivalent to the relation $\widehat{N}_{0,Q}=0$.
3.3. The operators $\widehat{Z}_{2,Q}$, $\widehat{R}_{2,Q}$, and $\widehat{N}^0_{1,Q}$
Let $\widehat{\omega} \in \widehat{\mathfrak{N}}$, and let $\widehat{\psi}_Q=\widehat{\psi}_Q(\widehat{\omega}) \in \operatorname{Dom} \widehat{X}_0$ be a (weak) solution of the problem
Clearly, the right-hand side of this equation belongs to $\widehat{\mathfrak{N}}^\perp=\operatorname{Ran}\widehat{X}_0^*$, and so the solvability condition is satisfied. We define the operator $\widehat{Z}_{2,Q}\colon \widehat{\mathfrak H}\to \widehat{\mathfrak H}$ by the relation $\widehat{Z}_{2,Q} \widehat{u}=\widehat{\psi}_Q(\widehat{P} \widehat{u})$, $\widehat{u} \in \widehat{\mathfrak H}$. Next, we define the operator $\widehat{R}_{2,Q}\colon \widehat{\mathfrak N} \to {\mathfrak H}_*$ by the relation $\widehat{R}_{2,Q}:=\widehat{X}_0 \widehat{Z}_{2,Q}+ \widehat{X}_1 \widehat{Z}_{Q}$.
Finally, we define the operator $\widehat{N}^0_{1,Q}$ by
3.4. The relationship between operators and the coefficients of power series expansions
Now we describe the relationship between the coefficients of the power series expansions (1.4) and (1.5) and the operators $\widehat{S}$ and $Q_{\widehat{\mathfrak{N}}}$ (see [8], §§ 1.6 and 1.7). We put $\zeta_l:=M \omega_l \in \widehat{\mathfrak{N}}$, $l=1,\dots,n$. Then from (1.7), (3.2), and (3.3) it follows that
The operators $\widehat{N}_{0,Q}$ and $\widehat{N}_{*,Q}$ can be described in terms of the coefficients of the power series expansions (1.4) and (1.5); cf. (1.11). Let $\widetilde{\zeta}_l:=M \widetilde{\omega}_l \in \widehat{\mathfrak{N}}$, $l=1,\dots,n$, where the elements $\widetilde{\omega}_l$ were defined by (1.6). Then
Relations (1.15) imply that $(Q_{\widehat{\mathfrak{N}}}\widetilde{\zeta}_j,\zeta_l)+ (\zeta_j,Q_{\widehat{\mathfrak{N}}}\widetilde{\zeta}_l)=0$, $j, l=1,\dots,n$. It follows that $(\widehat{N}_{*,Q}\zeta_j,\zeta_l)=0$ if $\gamma_j=\gamma_l$.
Now we return to the notation of § 1.6. Recall that the different eigenvalues of $S$ are denoted by $\gamma^{\circ}_j$, $j=1,\dots,p$, and $\mathfrak{N}_j=\operatorname{Ker}({S}-\gamma_j^\circ I_{\mathfrak{N}})$ are the corresponding eigenspaces. The vectors $\omega^{(j)}_i$, $i=1,\dots,k_j$, form an orthonormal basis in $\mathfrak{N}_j$. Then the same numbers $\gamma^{\circ}_j$, $j=1,\dots,p$, are the different eigenvalues of the problem (3.12), and $M \mathfrak{N}_j=\operatorname{Ker}(\widehat{S}- \gamma_j^\circ Q_{\widehat{\mathfrak{N}}})=:\widehat{\mathfrak{N}}_{j,Q}$ are the corresponding eigenspaces. The vectors $\zeta^{(j)}_i=M\omega^{(j)}_i$, $i=1,\dots,k_j$, form a basis in $\widehat{\mathfrak{N}}_{j,Q}$ which is orthonormal with weight $Q_{\widehat{\mathfrak{N}}}$. By $\mathcal{P}_j$ we denote the ‘skew’ projection onto $\widehat{\mathfrak{N}}_{j,Q}$ which is orthogonal with respect to the inner product $(Q_{\widehat{\mathfrak{N}}}\,{\cdot}\,{,}\,{\cdot}\,)$, that is,
It is easily seen that $\mathcal{P}_j=M P_j M^{-1} \widehat{P}$. There are analogues of relations (1.18). Using (1.18), (3.6), and (3.10) it is easy to obtain the invariant representations
The coefficients $\nu_l$ in the expansions (1.4) and the vectors $\zeta_l=M \omega_l$, $l=1,\dots,n$, are the eigenvalues and eigenvectors of some problem; see [30], § 3.4. We need to describe this problem in the case where $\mu_l=0$, $l=1,\dots,n,$ that is, $\widehat{N}_{0,Q}=0$ (see also [32], Proposition 5.3).
Proposition 3.3 ([30]). Let $\widehat{N}_{0,Q}=0$. Suppose that the operator $\widehat{N}_{1,Q}^0$ is defined by (3.11). Let $\gamma_1^\circ,\dots,\gamma_p^\circ$ be the different eigenvalues of problem (3.12), and $k_1,\dots,k_p$ be their multiplicities. Let $\widehat{\mathfrak{N}}_{q,Q}=\operatorname{Ker}(\widehat{S}- \gamma_q^\circ Q_{\widehat{\mathfrak{N}}})$, and let $\widehat{P}_{q,Q}$ be the orthogonal projection of the space $\widehat{{\mathfrak H}}$ onto the subspace $\widehat{\mathfrak{N}}_{q,Q}$, $q=1,\dots,p$. We introduce operators $\widehat{\mathcal{N}}_Q^{(q)}$, $q=1,\dots,p$ : each operator $\widehat{\mathcal{N}}_Q^{(q)}$ acts on $\widehat{\mathfrak{N}}_{q,Q}$ and is given by the expression
Denote $i(q)=k_1+\dots+k_{q-1}+1$. Let $\nu_l$, $l=1,\dots,n,$ be the coefficients of $t^4$ in expansions (1.4), and let $\omega_l$ be the ‘embryos’ from the expansions (1.5). Let $\zeta_l=M \omega_l$, $l=1,\dots,n$. Denote ${Q}_{\widehat{\mathfrak{N}}_{q,Q}}= \widehat{P}_{q,Q} Q\big|_{\widehat{\mathfrak{N}}_{q,Q}}$. Then
4. Approximation for the sandwiched operator exponential of the operator $A(t)=M^*\widehat{A}(t)M$
4.1. Approximation for the sandwiched exponential in the operator norm on $\widehat{\mathfrak H}$
Suppose that the assumptions of § 3.1 are satisfied. We describe approximation for the exponential $e^{-i \varepsilon^{-2}\tau A(t)}$, where $A(t)$ is the family of the form (3.1), in terms of the germ $\widehat{S}$ of the operator $\widehat{A}(t)$ and the isomorphism $M$. It turns out that it is convenient to border the operator exponential by the factors $M$ and $M^{-1}$.
Let $M_0:=(Q_{\widehat{\mathfrak{N}}})^{-1/2}$. According to [109], (6.2), we have
Similarly, Theorems 2.2 and 2.3, by using Lemma 3.1 and inequality (4.3), yield the following statements; see [30], Theorems 3.5 and 3.6. Recall that the operator $\widehat{N}_Q$ is defined by (3.5), and the operator $\widehat{N}_{0,Q}$ is defined by (3.15).
Theorem 4.2 ([30]). Let $\widehat{N}_Q=0$. Then for $\varepsilon >0$, $\tau \in \mathbb{R}$, and $ |t| \leqslant t_0$ we have
Under certain additional assumptions this result was improved in [109], Theorems 6.7 and 6.8. The following two results are deduced from Theorems 2.5 and 2.6.
Theorem 4.5 ([109]). Suppose that $\widehat{N}_Q =0$. Then for $\varepsilon >0$, $\tau \in \mathbb{R}$, and $|t| \leqslant t_0$ we have
holds for all sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [29], Theorem 5.10, statement $2^\circ$ in [109], Theorem 7.3, and statement $3^\circ$ in [109], Theorem 7.5.
Further, Theorems 4.2, 4.3, 4.5, 4.6, 4.8, and 4.9 (about improvements of general results under certain additional assumptions), in their turn, are sharp.
Theorem 4.11 ([30], [109]). Suppose that $\widehat{N}_{0,Q}=0$ and $\widehat{\mathcal N}_Q^{(q)} \ne 0$ for some $q\in \{1,\dots,p\}$.
$1^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 2$. Then there does not exist a constant $C(\tau)$ such that estimate (4.4) holds for all sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 4$. Then there does not exist a constant $C(\tau)$ such that estimate (4.5) holds for all sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $\tau \ne 0$ and $0 \leqslant s < 3$. Then there does not exist a constant $C(\tau)$ such that estimate (4.6) holds for all sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [30], Theorem 3.9, statement $2^\circ$ in [109], Theorem 7.4, and statement $3^\circ$ in [109], Theorem 7.6.
4.5. Confirming the sharpness of results with respect to time
The following theorem confirms that in the general case Theorems 4.1, 4.4, and 4.7 are sharp with respect to the dependence of estimates on $\tau$ (for large $|\tau|$).
Theorem 4.12 ([30], [109]). Suppose that $\widehat{N}_{0,Q} \ne 0$.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau| =0$ and estimate (4.4) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ \tau^2 =0$ and estimate (4.5) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau| =0$ and estimate (4.6) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [30], Theorem 3.10, statement $2^\circ$ in [109], Theorem 7.9, and statement $3^\circ$ in [109], Theorem 7.11.
Further, Theorems 4.2, 4.3, 4.5, 4.6, 4.8, and 4.9 (about improvements of general results under certain additional assumptions), in their turn, are sharp.
Theorem 4.13 ([30], [109]). Suppose that $\widehat{N}_{0,Q}=0$ and $\widehat{\mathcal N}_Q^{(q)} \ne 0$ for some $q\in \{1,\dots,p\}$.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau|^{1/2}=0$ and estimate (4.4) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau| =0$ and estimate (4.5) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $C(\tau)$ such that $\lim_{\tau \to \infty} C(\tau)/ |\tau|^{1/2} =0$ and estimate (4.6) holds for all $\tau \in \mathbb{R}$ and sufficiently small $|t|$ and $\varepsilon > 0$.
Statement $1^\circ$ was proved in [30], Theorem 3.11, statement $2^\circ$ in [109], Theorem 7.10, and statement $3^\circ$ in [109], Theorem 7.12.
Chapter 2. Periodic differential operators in $L_2(\mathbb{R}^d; \mathbb{C}^n)$
5. The class of differential operators in $L_2(\mathbb{R}^d; \mathbb{C}^n)$
5.1. Lattices. Fourier series
Let $\Gamma$ be a lattice in $\mathbb{R}^d$ generated by the basis $\mathbf{a}_1,\dots,\mathbf{a}_d$, that is,
The basis $\mathbf{b}_1,\dots, \mathbf{b}_d$ dual to the basis $\mathbf{a}_1,\dots,\mathbf{a}_d$ is defined by the relations $\langle\mathbf{b}_l,\mathbf{a}_j\rangle=2 \pi \delta_{lj}$. This basis generates the lattice $\widetilde\Gamma$ dual to the lattice $\Gamma$. By $\widetilde\Omega$ we denote the central Brillouin zone of $\widetilde\Gamma$:
Denote $|\Omega|=\operatorname{meas}\Omega$ and $|\widetilde\Omega|=\operatorname{meas}\widetilde\Omega$ and note that $|\Omega|\,|\widetilde\Omega|=(2\pi)^d$. Let $r_0$ be the radius of the ball inscribed in $\operatorname{clos}\widetilde\Omega$, and let $r_1:=\max_{\mathbf{k} \in \partial\widetilde{\Omega}}|\mathbf{k}|$. Note that
Let $\widetilde H^1(\Omega;\mathbb{C}^n)$ be the subspace of functions from $H^1(\Omega;\mathbb{C}^n)$ whose $\Gamma$-periodic extension to $\mathbb{R}^d$ belongs to $H^1_{\rm loc}(\mathbb{R}^d; \mathbb{C}^n)$. We have
and the convergence of the series on the right-hand side is equivalent to the relation $\mathbf{v} \in \widetilde{H}^1(\Omega;\mathbb{C}^n)$. From (5.1), (5.4), and (5.5) it follows that
First, we define the Gelfand transform $\mathcal{U}$ for functions in the Schwartz class $\mathbf{v} \in \mathcal{S}(\mathbb{R}^d;\mathbb{C}^n)$ by the formula
We have $\|\widetilde{\mathbf{v}}\|_{L_2(\widetilde{\Omega} \times \Omega)}= \|{\mathbf v}\|_{L_2(\mathbb{R}^d)}$, and $\mathcal{U}$ extends by continuity to a unitary mapping
The relation $\mathbf{v} \in H^1(\mathbb{R}^d;\mathbb{C}^n)$ is equivalent to the fact that $\widetilde{\mathbf{v}}(\mathbf{k},\,{\cdot}\,) \in \widetilde H^1(\Omega;\mathbb{C}^n)$ for almost all $\mathbf{k} \in \widetilde\Omega $ and
Under the transform $\mathcal{U}$, the operator of multiplication by a bounded $\Gamma$-periodic function in $L_2(\mathbb{R}^d;\mathbb{C}^n)$ turns to multiplication by the same function on fibres of the direct integral $\mathcal{H}$. The action of the first-order differential operator $b(\mathbf{D})$ on $\mathbf{v} \in H^1(\mathbb{R}^d;\mathbb{C}^n)$ turns to the action of the operator $b(\mathbf{D}+\mathbf{k})$ on $\widetilde{\mathbf{v}}(\mathbf{k},\,{\cdot}\,) \in \widetilde H^1(\Omega;\mathbb{C}^n)$.
Let $b (\mathbf{D})= \sum_{l=1}^d b_l D_l$, where the $b_l$ are constant $m \times n$ matrices (in general, with complex entries). Suppose that $m \geqslant n$. Consider the symbol $b(\boldsymbol{\xi})= \sum_{l=1}^d b_l \xi_l$, and suppose that $\operatorname{rank}b(\boldsymbol{\xi})=n$ for $0 \ne \boldsymbol{\xi} \in \mathbb{R}^d$. This is equivalent to the inequalities
Suppose that $f(\mathbf{x})$, $\mathbf{x} \in \mathbb{R}^d$, is a $\Gamma$-periodic $n \times n$ matrix-valued function and $h(\mathbf{x})$, $\mathbf{x} \in \mathbb{R}^d$, is a $\Gamma$-periodic $m \times m$ matrix-valued function. Assume that
The operator $\mathcal{X}$ is closed. The selfadjoint operator $\mathcal{A}=\mathcal{X}^*\mathcal{X}$ in $L_2(\mathbb{R}^d;\mathbb{C}^n)$ is generated by the closed quadratic form $\mathfrak{a}[\mathbf{u},\mathbf{u}]= \|\mathcal{X}\mathbf{u}\|^2_{L_2(\mathbb{R}^d)}$, $\mathbf{u} \in \operatorname{Dom}\mathcal{X}$. Formally,
A selfadjoint operator $\mathcal{A}(\mathbf{k})= \mathcal{X}(\mathbf{k})^*\mathcal{X}(\mathbf{k}) \colon \mathfrak{H} \to \mathfrak{H}$ is generated by the closed quadratic form $\mathfrak{a}(\mathbf{k})[\mathbf{u}, \mathbf{u}]= \|\mathcal{X}(\mathbf{k})\mathbf{u}\|_{\mathfrak{H}_*}^2$, $\mathbf{u} \in \mathfrak{d}$. Formally, we can write
As follows from (5.2) and (5.5) for $\mathbf{k}=0$, a function $\mathbf{v} \in \widetilde{H}^1(\Omega;\mathbb{C}^n)$ such that $ \int_{\Omega}\mathbf{v} \, d \mathbf{x}=0$ (that is, $\widehat{\mathbf{v}}_0=0$) satisfies
From (5.18) and the lower estimate (5.13) for $\mathbf{k}=0$ it follows that the distance $d^0$ of the point $\lambda_0=0$ to the rest of the spectrum of the operator $\mathcal{A}(0)$ satisfies the estimate
Denote by $E_j(\mathbf{k})$, $j \in \mathbb{N}$, the consecutive eigenvalues (counting multiplicities) of the operator $\mathcal{A}(\mathbf{k})$ (band functions):
The band functions $E_j(\mathbf{k})$ are continuous and $\widetilde{\Gamma}$-periodic. As shown in [7], Chap. 2, § 2.2 (using variational arguments), band functions satisfy the estimates
Conversely, if $\widetilde{\mathbf{u}} \in \mathcal{H}$ satisfies (5.22) and the integral in (5.23) is finite, then $\mathbf{u} \in \operatorname{Dom}\mathcal{X}$ and (5.23) is valid.
From (5.21) it follows that the spectrum of $\mathcal{A}$ coincides with the union of the bands $\operatorname{Ran}E_j$, $j \in\mathbb{N}$. By (5.16) and (5.17) we have $\min_{\mathbf{k}}E_j(\mathbf{k})=E_j(0)=0$ for $j=1,\dots,n$, so that the first $n$ spectral bands of the operator $\mathcal{A}$ overlap and have the common bottom $\lambda_0=0$, while the $(n+1)$st band is separated from zero (see (5.20)).
5.7. Incorporating the operators $\mathcal{A}(\mathbf{k})$ in the abstract scheme
If $d > 1$, then the operators $\mathcal{A}(\mathbf{k})$ depend on the multidimensional parameter $\mathbf{k}$. In accordance with [7], Chap. 2, we introduce the one-dimensional parameter $t=|\mathbf{k}|$. We are based on the scheme from Chapter 1. Now all constructions depend on the additional parameter $\boldsymbol{\theta}=\mathbf{k}/|\mathbf{k}| \in \mathbb{S}^{d-1}$, and we have to make estimates uniform in $\boldsymbol{\theta}$. The spaces $\mathfrak{H}$ and $\mathfrak{H}_*$ are defined by (5.12). We put $X(t)=X(t,\boldsymbol{\theta}):=\mathcal{X}(t \boldsymbol{\theta})$. Then $X(t,\boldsymbol{\theta})= X_0+t X_1(\boldsymbol{\theta})$, where $X_0=h(\mathbf{x})b(\mathbf{D})f(\mathbf{x})$, $\operatorname{Dom}X_0=\mathfrak{d}$, and $X_1(\boldsymbol{\theta})$ is a bounded operator of multiplication by the matrix $h(\mathbf{x}) b(\boldsymbol{\theta}) f(\mathbf{x})$. Next, we put $A(t)=A(t,\boldsymbol{\theta}):=\mathcal{A}(t \boldsymbol{\theta})$. The kernel $\mathfrak{N}=\operatorname{Ker}X_0=\operatorname{Ker}\mathcal{A}(0)$ is defined by (5.17), $\dim \mathfrak{N}=n$. The number $d^0$ satisfies estimate (5.19). As shown in [7], Chap. 2, § 3, the condition $n \leqslant n_*=\dim\operatorname{Ker}X^*_0$ is also satisfied. Moreover, either $n_*=n$ (if $m=n$), or $n_*=\infty$ (if $m > n$). Thus, all the assumptions of the abstract scheme are satisfied.
According to § 1.1, we must fix a number $\delta >0$ such that $\delta < d^0/8$. Using (5.15) and (5.19) we put
Obviously, $t_0 \leqslant r_0/2$. Hence the ball $|\mathbf{k}| \leqslant t_0$ lies fully in $\widetilde{\Omega}$. It is important that $c_*$, $\delta$, and $t_0$ (see (5.15), (5.24), and (5.26)) do not depend on $\boldsymbol{\theta}$.
By (5.14) Condition 1.4 is satisfied. The germ $S(\boldsymbol{\theta})$ of the operator $A(t,\boldsymbol{\theta})$ is non-degenerate uniformly in $\boldsymbol{\theta}$ (cf. (1.17)):
6. Effective characteristics of the operator $\widehat{\mathcal{A}}$
6.1. The operator $A(t,\boldsymbol{\theta})$ in the case where $f=\mathbf{1}_n$
A special role is played by the operator $\mathcal A$ for $f=\mathbf{1}_n$. In this case we agree to mark all objects by hats ‘$\widehat{\phantom{\_}} $’. Then for the operator
the family $\widehat{\mathcal{A}}(\mathbf{k})=b(\mathbf{D}+\mathbf{k})^* g(\mathbf{x}) b(\mathbf{D}+\mathbf{k})$ is denoted by $\widehat{A}(t,\boldsymbol{\theta})$. The kernel (5.17) takes the form
so that $\widehat{\mathfrak{N}}$ consists of constant vector-valued functions. The orthogonal projection $\widehat{P}$ of the space $L_2(\Omega;\mathbb{C}^n)$ onto the subspace (6.2) is the operator of averaging over the cell:
Now the operators $\widehat{Z}(\boldsymbol{\theta})$, $\widehat{R}(\boldsymbol{\theta})$, and $\widehat{S}(\boldsymbol{\theta})$ for the family $\widehat{A}(t,\boldsymbol{\theta})$ (defined in § 1.2 in abstract terms) depend on $\boldsymbol{\theta}$. They were found in [9], § 4.1, and [7], Chap. 3, § 1.
Let $\Lambda \in \widetilde{H}^1(\Omega)$ be a $\Gamma$-periodic $n \times m$ matrix-valued function satisfying the equation
Then the operators $\widehat{Z}(\boldsymbol{\theta})\colon \mathfrak{H} \to \mathfrak{H}$ and $\widehat{R}(\boldsymbol{\theta})\colon\widehat{\mathfrak{N}} \to \mathfrak{N}_*$ are represented as
Here and in what follows square brackets denote the operator of multiplication by a function. The spectral germ $\widehat{S} (\boldsymbol{\theta})= \widehat{R}(\boldsymbol{\theta})^*\widehat{R}(\boldsymbol{\theta})$ of the family $\widehat{A}(t,\boldsymbol{\theta})$ acting on $\widehat{\mathfrak{N}}$ is given by
where $b(\boldsymbol{\theta})$ is the symbol of the operator $b(\mathbf{D})$ and $g^0$ is the so-called effective matrix. It is defined in terms of the matrix $\Lambda(\mathbf{x})$:
which acts in $L_2(\mathbb{R}^d;\mathbb{C}^n)$ and is called the effective operator for the operator $\widehat{\mathcal{A}}$.
Let $\widehat{\mathcal{A}}^{\,0} (\mathbf{k})$ be the operator family in $L_2(\Omega; \mathbb{C}^n)$ corresponding to the operator (6.17). Then $\widehat{\mathcal{A}}^{\,0}(\mathbf{k})= b(\mathbf{D}+\mathbf{k})^*g^0 b(\mathbf{D}+\mathbf{k})$ with periodic boundary conditions. In combination with (6.3) and (6.16) this implies that
$$
\begin{equation*}
\overline{g}:=|\Omega|^{-1}\int_{\Omega}g(\mathbf{x})\, d \mathbf{x} \quad\textit{and}\quad \underline{g}:=\biggl(|\Omega|^{-1} \int_{\Omega}g(\mathbf{x})^{-1}\, d \mathbf{x}\biggr)^{-1}.
\end{equation*}
\notag
$$
If $m=n$, then $g^0=\underline{g}$.
Estimates (6.19) are known in homogenization theory for particular differential operators as the Voigt–Reuss bracketing. Note also that estimates (6.19) imply that
where the $\mathbf{l}_k(\mathbf{x})$, $k=1,\dots,m,$ are the columns of the matrix $g (\mathbf{x})^{-1}$.
6.5. Analytic branches of the eigenvalues and eigenvectors
The analytic (in $t$) branches of the eigenvalues $\widehat{\lambda}_l(t,\boldsymbol{\theta})$ and the eigenvectors $\widehat{\varphi}_l(t,\boldsymbol{\theta})$ of the operator $\widehat{A}(t,\boldsymbol{\theta})$ admit power series expansions of the form (1.4) and (1.5), with coefficients depending on $\boldsymbol{\theta}$ (we do not control the interval of convergence $t=|\mathbf{k}| \leqslant t_*(\boldsymbol{\theta})$):
According to (1.7), the numbers $\widehat{\gamma}_l(\boldsymbol{\theta})$ and the elements $\widehat{\omega}_l(\boldsymbol{\theta})$ are eigenvalues and eigenvectors of the germ:
6.6. The operator $\widehat{N}(\boldsymbol{\theta})$
Let us describe the operator $N$ (defined by (1.13) in abstract terms). As shown in [9], § 4, for the family $\widehat{A}(t,\boldsymbol{\theta})$ this operator takes the form
where $L (\boldsymbol{\theta})$ is the $m \times m$ matrix-valued function given by
$$
\begin{equation}
L (\boldsymbol{\theta})=|\Omega|^{-1}\int_{\Omega}\bigl(\Lambda(\mathbf{x})^* b(\boldsymbol{\theta})^* \widetilde{g}(\mathbf{x})+ \widetilde{g}(\mathbf{x})^* b(\boldsymbol{\theta}) \Lambda(\mathbf{x})\bigr)\, d \mathbf{x}.
\end{equation}
\tag{6.25}
$$
Here $\Lambda(\mathbf{x})$ is the $\Gamma$-periodic solution of problem (6.8) and $\widetilde{g}(\mathbf{x})$ is the matrix-valued function (6.11).
Note that $L(\mathbf{k}):=tL(\boldsymbol{\theta}),\mathbf{k} \in \mathbb{R}^d$, is a Hermitian first-order homogeneous matrix- valued function. We put $\widehat{N}(\mathbf{k}):= t^3 \widehat{N} (\boldsymbol{\theta})$, $\mathbf{k} \in \mathbb{R}^d$. Then
The matrix-valued function $b(\mathbf{k})^*L(\mathbf{k})b(\mathbf{k})$ is a third-order homogeneous polynomial of $\mathbf{k} \in \mathbb{R}^d$.
In [9] some conditions sufficient for the equality $\widehat{N}(\boldsymbol{\theta})=0$ were presented.
Proposition 6.4 ([9], § 4). Suppose that at least one of the following assumptions is satisfied :
(a) The operator $\widehat{\mathcal{A}}$ is given by $\widehat{\mathcal{A}}=\mathbf{D}^* g(\mathbf{x})\mathbf{D}$, where $g(\mathbf{x})$ is a symmetric matrix with real entries;
(b) Relations (6.20) are satisfied, that is, $g^0=\overline{g}$;
(c) Relations (6.21) are satisfied, that is, $g^0=\underline{g}$ (in particular, they are valid if $m=n$).
Then $\widehat{N} (\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
On the other hand, in [9], §§ 10.4, 13.2, and 14.6, there are examples of operators $\widehat{\mathcal{A}}$ for which the operator $\widehat{N}(\boldsymbol{\theta})$ is not equal to zero. This is an example of a scalar elliptic operator (the case where $n=1$) with Hermitian matrix of coefficients with complex entries, and also an example of a matrix operator with real-valued coefficients (see also [29], Example 8.7, and [31], § 14.3). Recall (see Remark 1.3) that $\widehat{N}(\boldsymbol{\theta})=\widehat{N}_0 (\boldsymbol{\theta})+ \widehat{N}_*(\boldsymbol{\theta})$, where the operator $\widehat{N}_0(\boldsymbol{\theta})$ is diagonal in the basis $\{\widehat{\omega}_l(\boldsymbol{\theta})\}_{l=1}^n$, and the operator $\widehat{N}_*(\boldsymbol{\theta})$ has zero diagonal entries. We have
Proposition 6.5. Suppose that the matrices $b(\boldsymbol{\theta})$ and $g(\mathbf{x})$ have real entries. Suppose that in the expansions (6.23) for analytic branches of eigenvectors of the operator $\widehat{A}(t,\boldsymbol{\theta})$ the ‘embryos’ $\widehat{\omega}_l(\boldsymbol{\theta})$, $l=1,\dots,n$, can be chosen real. Then the coefficients $\widehat{\mu}_l(\boldsymbol{\theta})$, $l=1,\dots,n$, in (6.22) are equal to zero, that is, $\widehat{N}_0(\boldsymbol{\theta})=0$.
In the ‘real’ case under consideration the germ $\widehat{S}(\boldsymbol{\theta})$ is a symmetric matrix with real entries. Clearly, in the case of a simple eigenvalue $\widehat{\gamma}_j(\boldsymbol{\theta})$ of the germ, the vector $\widehat{\omega}_j(\boldsymbol{\theta})$ is determined uniquely up to a phase factor, and it can always be chosen real. We arrive at the following result.
Corollary 6.6. Suppose that the matrices $b(\boldsymbol{\theta})$ and $g(\mathbf{x})$ have real entries. Suppose also that the spectrum of the germ $\widehat{S}(\boldsymbol{\theta})$ is simple. Then $\widehat{N}_0(\boldsymbol{\theta})=0$.
However, according to examples in [29] and [31] (see [29], Example 8.7, and [31], § 14.3), in the ‘real’ case it is not always possible to choose the vectors $\widehat{\omega}_l(\boldsymbol{\theta})$ to be real. It can happen that $\widehat{N}_0(\boldsymbol{\theta}) \ne 0$ at some points $\boldsymbol{\theta}$.
6.7. The operators $\widehat{Z}_2(\boldsymbol{\theta})$, $\widehat{R}_2(\boldsymbol{\theta})$, and $\widehat{N}_1^0(\boldsymbol{\theta})$
We describe the operators $Z_2$, $R_2$, and $N_1^0$ (in abstract terms they were defined in §§ 1.2 and 1.7) for the family $\widehat{A}(t,\boldsymbol{\theta})$. Suppose that an $ n \times m $ matrix-valued function $\Lambda_l^{(2)}({\mathbf x})$ is the $\Gamma$-periodic solution of the problem
6.8. The multiplicities of eigenvalues of the germ
In this subsection we assume that $n \geqslant 2$. We go over to the notation adopted in § 1.6, keeping track of the multiplicities of eigenvalues of the spectral germ $\widehat{S}(\boldsymbol{\theta})$. In general, the number $p(\boldsymbol{\theta})$ of the different eigenvalues $\widehat{\gamma}^{\circ}_1(\boldsymbol{\theta}),\dots, \widehat{\gamma}^{\circ}_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ of the spectral germ $\widehat{S}(\boldsymbol{\theta})$ and their multiplicities $k_1(\boldsymbol{\theta}),\dots, k_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ depend on the parameter $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. For each fixed $\boldsymbol{\theta}$, let $\widehat{P}_j(\boldsymbol{\theta})$ be the orthogonal projection of $L_2(\Omega;\mathbb{C}^n)$ onto the subspace $\widehat{\mathfrak N}_j(\boldsymbol{\theta})$ of the germ $\widehat{S}(\boldsymbol{\theta})$ corresponding to the eigenvalue $\widehat{\gamma}_j^{\circ}(\boldsymbol{\theta})$. We have the following invariant representations for the operators $\widehat{N}_0(\boldsymbol{\theta})$ and $\widehat{N}_*(\boldsymbol{\theta})$:
6.9. The coefficients $\widehat{\nu}_l(\boldsymbol{\theta})$
The coefficients $\widehat{\nu}_l(\boldsymbol{\theta})$, $l=1,\dots,n$, in expansions (6.22) are the eigenvalues of some problem. We need to describe this problem in the case where $\widehat{\mu}_l(\boldsymbol{\theta})=0$, $l=1,\dots,n$, that is, $\widehat{N}_0(\boldsymbol{\theta})=0$. Applying Proposition 1.6 we arrive at the following statement (see also [32], Proposition 8.7).
Proposition 6.7. Let $\widehat{N}_0(\boldsymbol{\theta})=0$. Suppose that $\widehat{\gamma}_1^\circ(\boldsymbol{\theta}),\dots, \widehat{\gamma}_{p(\boldsymbol{\theta})}^\circ(\boldsymbol{\theta})$ are the different eigenvalues of the operator (6.10), and $k_1(\boldsymbol{\theta}),\dots, k_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ are their multiplicities. Let $\widehat{P}_q(\boldsymbol{\theta})$ be the orthogonal projection of the space $L_2(\Omega;\mathbb{C}^n)$ onto the subspace $\widehat{\mathfrak{N}}_q (\boldsymbol{\theta}) = \operatorname{Ker}\bigl(\widehat{S}(\boldsymbol{\theta}) -\widehat{\gamma}_q^\circ(\boldsymbol{\theta}) I_{\widehat{\mathfrak{N}}}\bigr)$, $q=1,\dots,p(\boldsymbol{\theta})$. Let $\widehat{Z}(\boldsymbol{\theta})$ and $\widehat{N}_1^0(\boldsymbol{\theta})$ be the operators defined by (6.9) and (6.27), (6.28), respectively. We introduce the operators $\widehat{\mathcal{N}}^{(q)}(\boldsymbol{\theta})$, $q=1,\dots,p(\boldsymbol{\theta})$: the operator $\widehat{\mathcal{N}}^{(q)}(\boldsymbol{\theta})$ acts on $\widehat{\mathfrak{N}}_q(\boldsymbol{\theta})$ and is given by the expression
Set $i(q,\boldsymbol{\theta})=k_1(\boldsymbol{\theta})+\dots+ k_{q-1}(\boldsymbol{\theta})+1$. Let $\widehat{\nu}_l(\boldsymbol{\theta})$ be the coefficients of $t^4$ in expansions (6.22), and let $\widehat{\omega}_l(\boldsymbol{\theta})$ be the ‘embryos’ from (6.23), $l=1,\dots,n$. Then
7. Approximation of the operator $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal A}(\mathbf{k})}$
7.1. Approximation in the operator norm on $L_2(\Omega;\mathbb{C}^n)$
Consider the operator $\mathcal{H}_0=-\Delta$ in $L_2(\mathbb{R}^d;\mathbb{C}^n)$. In the direct integral expansion the operator $\mathcal{H}_0$ is associated with the family of operators $\mathcal{H}_0(\mathbf{k})$ acting on $L_2(\Omega;\mathbb{C}^n)$. The operator $\mathcal{H}_0(\mathbf{k})$ is given by the differential expression $|\mathbf{D}+\mathbf{k}|^2$ with periodic boundary conditions. Denote
We have taken into account that $|\mathbf{b}+\mathbf{k}| \geqslant r_0$ for $0 \ne \mathbf{b} \in \widetilde{\Gamma}$ and $\mathbf k \in \widetilde{\Omega}$ (see (5.1)).
We apply theorems from § 2 to the operator $\widehat{A}(t,\boldsymbol{\theta})= \widehat{\mathcal{A}}(\mathbf{k})$. According to Remark 2.10, we can track the dependence of the constants in estimates on the problem data. Note that $\widehat{c}_*$, $\widehat{\delta}$, and $\widehat{t}_0$ do not depend on $\boldsymbol{\theta}$ (see (6.4)–(6.6)). According to (6.7), the norm $\|\widehat{X}_1(\boldsymbol{\theta})\|$ can be replaced by $\alpha_1^{1/2}\|g\|_{L_{\infty}}^{1/2}$. Therefore, the constants in Theorem 2.1 (as applied to the operator $\widehat{\mathcal{A}}(\mathbf{k})$) do not depend on $\boldsymbol{\theta}$. They depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 7.1 ([25]). For $\tau \in \mathbb{R}$, $\varepsilon >0$, and $\mathbf k \in \widetilde{\Omega}$ we have
Previously, Theorem 7.1 was obtained in [25], Theorem 7.1.
Now we improve the result of Theorem 7.1 under some additional assumptions. We impose the following condition.
Condition 7.2. Let $\widehat{N}(\boldsymbol{\theta})$ be the operator defined by (6.24). Then suppose that $\widehat{N}(\boldsymbol{\theta})=0$ for all $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
The following result is deduced from Theorem 2.2; it was obtained in [30], Theorem 6.2.
Theorem 7.3 ([30]). Suppose that Condition 7.2 is fulfilled. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant $\widehat{\mathcal C}_2$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Now we abandon the assumption that $\widehat{N}(\boldsymbol{\theta}) \equiv 0$, but we assume instead that $\widehat{N}_0(\boldsymbol{\theta})=0$ for all $\boldsymbol{\theta}$. Assume that $\widehat{N}(\boldsymbol{\theta})= \widehat{N}_*(\boldsymbol{\theta})\ne 0$ for some $\boldsymbol{\theta}$ (otherwise Theorem 7.3 applies). We would like to use an ‘abstract’ result (namely, Theorem 2.3). However, there is an additional complication associated with the fact that the multiplicity of the spectrum of the germ $\widehat{S}(\boldsymbol{\theta})$ can change at some points $\boldsymbol{\theta}$. When approaching such points, the distance between some pair of different eigenvalues of the germ tends to zero, and we cannot choose the quantities $\widehat{c}^{\circ}_{jl}$ and $\widehat{t}^{00}_{jl}$ to be independent of $\boldsymbol{\theta}$. Therefore, we are forced to impose additional conditions. We need to take care only about those eigenvalues for which the corresponding term in representation (6.30) is non-trivial. Since the number of different eigenvalues of the germ and their multiplicities can depend on $\boldsymbol{\theta}$, to formulate the additional condition it is more convenient to use the initial numbering of the eigenvalues $\widehat{\gamma}_1(\boldsymbol{\theta}),\dots, \widehat{\gamma}_n(\boldsymbol{\theta})$ of the germ $\widehat{S}(\boldsymbol{\theta})$ (each eigenvalue repeated as many times as its multiplicity is), having agreed to number them in the non-decreasing order:
For each $\boldsymbol{\theta}$ we denote by $\widehat{P}^{(k)}(\boldsymbol{\theta})$ the orthogonal projection of the space $L_2(\Omega;\mathbb{C}^n)$ onto the eigenspace of the operator $\widehat{S}(\boldsymbol{\theta})$ corresponding to the eigenvalue $\widehat{\gamma}_k(\boldsymbol{\theta})$. Clearly, for every $\boldsymbol{\theta}$ the operator $\widehat{P}^{(k)}(\boldsymbol{\theta})$ coincides with one of the projections $\widehat{P}_j(\boldsymbol{\theta})$ introduced in § 6.8 (but the index $j$ can depend on $\boldsymbol{\theta}$ and changes at points where the multiplicity of the spectrum of the germ changes).
Condition 7.4. $1^\circ$. The operator $\widehat{N}_0(\boldsymbol{\theta})$ defined by (6.29) is equal to zero: $\widehat{N}_0(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
$2^\circ$. For each pair of indices $(k,r)$, $1 \leqslant k,r \leqslant n$, $k \ne r$, such that $\widehat{\gamma}_k(\boldsymbol{\theta}_0)= \widehat{\gamma}_r(\boldsymbol{\theta}_0)$, for some $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$ we have $\widehat{P}^{(k)}(\boldsymbol{\theta})\widehat{N}(\boldsymbol{\theta}) \widehat{P}^{(r)}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Condition $2^\circ$ can be reformulated: we require that, for ‘blocks’ $\widehat{P}^{(k)}(\boldsymbol{\theta})\widehat{N}(\boldsymbol{\theta}) \widehat{P}^{(r)}(\boldsymbol{\theta})$ of the operator $\widehat{N}(\boldsymbol{\theta})$ which are distinct from identical zero, the corresponding branches of eigenvalues $\widehat{\gamma}_k(\boldsymbol{\theta})$ and $\widehat{\gamma}_r(\boldsymbol{\theta})$ do not intersect. Of course, Condition 7.4 is ensured by the following stronger condition.
Condition 7.5. $1^\circ$. The operator $\widehat{N}_0(\boldsymbol{\theta})$ defined by (6.29) is equal to zero: $\widehat{N}_0(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
$2^\circ$. The number $p$ of different eigenvalues of the spectral germ $\widehat{S}(\boldsymbol{\theta})$ does not depend on $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Remark 7.6. The assumption $2^\circ$ of Condition 7.5 is a fortiori satisfied if the spectrum of the germ $\widehat{S}(\boldsymbol{\theta})$ is simple for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Thus, we assume that Condition 7.4 is satisfied. We are only interested in pairs of indices from the set
Since the operator $\widehat{S}(\boldsymbol{\theta})$ depends on $\boldsymbol{\theta}\in \mathbb{S}^{d-1}$ continuously (it is a polynomial of the second order), the perturbation theory of a discrete spectrum shows that the functions $\widehat{\gamma}_j(\boldsymbol{\theta})$ are continuous on the sphere $\mathbb{S}^{d-1}$. By assumption $2^\circ$ of Condition 7.4, for $(k,r) \in \widehat{\mathcal{K}}$ we have $|\widehat{\gamma}_k(\boldsymbol{\theta})- \widehat{\gamma}_r(\boldsymbol{\theta})| > 0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$, so that
Clearly, the number (7.8) is a realization of the quantity (2.1) chosen independent of $\boldsymbol{\theta}$. Under Condition 7.4, the number subject to (2.2) can also be chosen independent of $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. Taking (6.5) and (6.7) into account, we put
where $\widehat{c}^{\circ}$ is defined by (7.8). (The condition $\widehat{t}^{\,00} \leqslant \widehat{t}_{0}$ is valid since $\widehat{c}^{\circ}\! \leqslant \|\widehat{S}(\boldsymbol{\theta})\| \leqslant \alpha_1\|g\|_{L_{\infty}}$.)
Remark 7.7. In contrast to the number $\widehat{t}_{0}$ (see (6.6)), which is controlled only via $r_0$, $\alpha_0$, $\alpha_1$, $\|g\|_{L_{\infty}}$, and $\|g^{-1}\|_{L_{\infty}}$, the value $\widehat{t}^{\,00}$ depends on the spectral characteristics of the germ, namely, the minimum distance between its different eigenvalues $\widehat{\gamma}_k(\boldsymbol{\theta})$ and $ \widehat{\gamma}_r (\boldsymbol{\theta})$ (where $(k,r)$ runs through $\widehat{\mathcal{K}}$).
Under Condition 7.4 we deduce the following result from Theorem 2.3 (see [30], Theorem 6.7). Now the constants in estimates depend not only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$, but also on $\widehat{c}^\circ$ and $n$; see Remark 2.10.
Theorem 7.8 ([30]). Let Condition 7.4 (or the more restrictive Condition 7.5) be fulfilled. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant $\widehat{\mathcal C}_3$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^\circ$.
7.2. More accurate approximation in the operator norm on $L_2(\Omega;\mathbb{C}^n)$
Note that the operator (7.14) is bounded and the operator (7.15) is in the general case defined on $\widetilde{H}^3({\Omega};\mathbb{C}^n)$. The operators (7.14) and (7.15) can be represented as
From (7.9), (7.10), (7.12), (7.13), and (7.16)–(7.19) it follows that for $\varepsilon >0$, $\tau \in \mathbb{R}$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
Theorem 7.9. Suppose that the operator $\widehat{G}(\mathbf{k},\varepsilon^{-2}\tau)$ is defined by (7.15). Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant $\widehat{\mathcal{C}}''_4$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Now we show that the operator $\widehat{G}(\mathbf{k},\varepsilon^{-2}\tau) \mathcal{R}(\mathbf{k},\varepsilon)^3 \widehat{P}$ in inequality (7.26) can be replaced by $\widehat{G}(\mathbf{k}, \varepsilon^{-2}\tau)\mathcal{R}(\mathbf{k},\varepsilon)^3$ (within the permissible error). To do this we estimate the operator
By (7.4), (7.6), and (7.7) the norm of the first term does not exceed $2 r_0^{-2}\varepsilon^2$. It is easy to estimate the second term using the discrete Fourier transform. Its norm does not exceed the quantity
Comparing estimates (7.26) and (7.27) we obtain the required inequality (7.23). $\Box$
Now, using Theorems 2.5 and 2.6, we improve the result of Theorem 7.9 under some additional assumptions.
Theorem 7.10. Suppose that the operator $\widehat{G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ is defined by (7.14). Let Condition 7.2 be fulfilled. Then for $\tau \in \mathbb{R}$, $\varepsilon>0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
By (7.4), (7.6), and (7.7) the norm of this operator is estimated by $2r_0^{-2}\varepsilon^2$. In combination with (7.29) and (7.30), this implies the required estimate (7.28). $\Box$
Theorem 7.11. Let $\widehat{G}(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (7.15). Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon>0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant $\widehat{\mathcal C}_6$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^\circ$.
The constant $\widehat{\mathcal{C}}'_6$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^\circ$.
For $|\mathbf{k}| > \widehat{t}^{\,00}$ we use (7.2) and (7.22):
Theorem 7.12. Suppose that the operator $\widehat{G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ is defined by (7.14). Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
Now, using Theorems 2.8 and 2.9, we improve the result of Theorem 7.12 under some additional assumptions.
Theorem 7.13. Suppose that the operator $\widehat{G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ is defined by (7.14). Let Condition 7.2 be fulfilled. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
Similarly, from Theorem 2.9 and Remark 2.10 we deduce the following result.
Theorem 7.14. Let $\widehat{G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (7.14). Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant $\widehat{\mathcal C}_{9}$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^\circ$.
8. Confirmation of the sharpness of the results on approximations of the operator $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal A}(\mathbf{k})}$
8.1. Sharpness with respect to the smoothing factor
In the statements of this section we impose one of the following two conditions.
Condition 8.1. Let $\widehat{N}_0(\boldsymbol{\theta})$ be the operator defined by (6.29). Then suppose that $\widehat{N}_0(\boldsymbol{\theta}_0) \ne 0$ at least at one point $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$.
Condition 8.2. Let $\widehat{N}_0(\boldsymbol{\theta})$ and $\widehat{\mathcal N}^{(q)}(\boldsymbol{\theta})$ be the operators defined by (6.29) and (6.31), respectively. Then suppose that $\widehat{N}_0(\boldsymbol{\theta})=0$ for all $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$ and
for some $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$ and some $q \in \{1,\dots, p(\boldsymbol{\theta}_0)\}$.
We need the following lemma (see [29], Lemma 9.9, and [32], Lemma 10.3).
Lemma 8.3 ([29], [32]). Let $\widehat{\delta}$ and $\widehat{t}_0$ be defined by (6.5) and (6.6), respectively. Let $\widehat{F}(\mathbf{k})$ be the spectral projection of the operator $\widehat{\mathcal{A}}(\mathbf{k})$ for the interval $[0,\widehat{\delta}]$. Then for $|\mathbf{k}| \leqslant \widehat{t}_0$ and $|\mathbf{k}_0| \leqslant \widehat{t}_0$ we have
holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
Proof. Statement $1^\circ$ was proved in [29], Theorem 9.8, on the basis of the abstract result of Theorem 2.11.
Let us prove $2^\circ$. It suffices to assume that $2 \leqslant s < 6$. We prove by contradiction. Suppose that for some $\tau \ne 0$ and $2 \leqslant s < 6$ there exists a constant $\mathcal{C}(\tau)$ such that estimate (8.2) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$. Multiplying the operator under the norm sign in (8.2) by $\widehat{P}$ and using (7.2) we see that the inequality
Note that the projection $\widehat{P}$ is the spectral projection of the operator $\widehat{\mathcal{A}}^{\,0}(\mathbf{k})$ for the interval $[0,\widehat{\delta}\,]$. Therefore, from Lemma 8.3 (as applied to $\widehat{\mathcal{A}}(\mathbf{k})$ and $\widehat{\mathcal{A}}^{\,0}(\mathbf{k})$) it follows that for fixed $\tau$ and $\varepsilon$ the operator $\widehat{\mathfrak G}(\mathbf{k},\varepsilon^{-2}\tau)$ is continuous with respect to $\mathbf{k}$ in the ball $|\mathbf{k}| \leqslant \widehat{t}_0$. Hence estimate (8.7) is valid for all values of $\mathbf{k}$ in this ball. In particular, it holds for $\mathbf{k}=t\boldsymbol{\theta}_0$ if $t \leqslant \widehat{t}_0$. Applying (8.6) once again, we see that for some constant $\widehat{\mathcal{C}}(\tau)$ the estimate
holds for all $t \leqslant \widehat{t}_0$ and sufficiently small $\varepsilon$.
Estimate (8.8) corresponds to the abstract estimate (2.4). Since, by Condition 8.1, $\widehat{N}_0(\boldsymbol{\theta}_0)\ne 0$, the assumptions of Theorem 2.11, $(2^\circ)$ are satisfied. Applying this theorem we arrive at a contradiction.
We proceed to the proof of statement $3^\circ$. It suffices to assume that $2 \leqslant s < 4$. We prove by contradiction. Suppose that for some $\tau \ne 0$ and $2 \leqslant s < 4$ there exists a constant $\mathcal{C}(\tau)$ such that estimate (8.3) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$. Multiplying the operator under the norm sign in (8.3) by $\widehat{P}$ and using (7.2), we see that the estimate
From Lemma 8.3 (as applied to $\widehat{\mathcal{A}} (\mathbf{k})$ and $\widehat{\mathcal{A}}^{\,0}(\mathbf{k})$) it follows that, for fixed $\tau$ and $\varepsilon$, the operator $\widehat{\mathcal A}(\mathbf k)^{1/2} {\mathfrak G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ is continuous with respect to $\mathbf{k}$ in the ball $|\mathbf{k}| \leqslant \widehat{t}_0$. Hence estimate (8.12) is valid for all values of $\mathbf{k}$ in this ball. In particular, it holds for $\mathbf{k}=t\boldsymbol{\theta}_0$ if $t \leqslant \widehat{t}_0$. Applying (8.11) once again, we see that for some constant $\widehat{\mathcal{C}}(\tau)$ the estimate
holds for all $t \leqslant \widehat{t}_0$ and sufficiently small $\varepsilon$.
Estimate (8.13) corresponds to the abstract estimate (2.5). Since, by Condition 8.1, $\widehat{N}_0 (\boldsymbol{\theta}_0)\ne 0$, the assumptions of Theorem 2.11($3^\circ$) are satisfied. Applying this theorem, we arrive at a contradiction. $\Box$
Similarly, from Theorem 2.12 we deduce the following result, which confirms the sharpness of Theorems 7.3, 7.8, 7.10, 7.11, 7.13, and 7.14 (about improvements of general results under additional assumptions).
Theorem 8.5. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C} (\tau)$ such that estimate (8.1) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C} (\tau)$ such that estimate (8.2) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C} (\tau)$ such that estimate (8.3) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
Statement $1^\circ$ was obtained in [30], Theorem 6.9.
8.2. Sharpness of results with respect to time
In this subsection we verify that the results of § 7 are sharp with respect to the dependence of estimates on $\tau$ (for large $|\tau|$).
By analogy with the proof of Theorem 8.4, from Theorem 2.13 we deduce the following result, which shows that Theorems 7.1, 7.9, and 7.12 are sharp. Statement $1^\circ$ was obtained in [30], Theorem 6.10.
Theorem 8.6. Suppose that Condition 8.1 is satisfied.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/|\tau|=0$ and estimate (8.1) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/\tau^2=0$ and estimate (8.2) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/|\tau|=0$ and estimate (8.3) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
In a similar way, from Theorem 2.14 we deduce the following result, which confirms the sharpness of Theorems 7.3, 7.8, 7.10, 7.11, 7.13, and 7.14. Statement $1^\circ$ was obtained in [30], Theorem 6.11.
Theorem 8.7. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau|^{1/2}=0$ and estimate (8.1) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau|=0$ and estimate (8.2) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau|^{1/2}=0$ and estimate (8.3) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
9. Effective characteristics of the operator $\mathcal{A}(\mathbf{k})$
9.1. Application of the scheme from § 3 to the operator $\mathcal{A}(\mathbf{k})$
In this section we study the operator $\mathcal{A}(\mathbf{k})=f^*\widehat{\mathcal{A}}(\mathbf{k})f$ following the scheme presented in § 3. Now we have $\mathfrak{H}=\widehat{\mathfrak{H}}=L_2(\Omega; \mathbb{C}^n)$ and $\mathfrak{H}_*=L_2(\Omega;\mathbb{C}^m)$. The role of the operator $A(t)$ is played by $A(t,\boldsymbol{\theta})=\mathcal{A}(\mathbf{k})$, and the role of the operator $\widehat{A}(t)$ is played by $\widehat{A}(t,\boldsymbol{\theta})= \widehat{\mathcal{A}}(\mathbf{k})$. The isomorphism $M$ is the operator of multiplication by the matrix-valued function $f(\mathbf{x})$. The operator $Q$ is the operator of multiplication by the matrix-valued function
9.2. Analytic branches of eigenvalues and eigenvectors
According to (3.3), the spectral germ $S(\boldsymbol{\theta})$ of the operator $A(t,\boldsymbol{\theta})$ acting on the subspace $\mathfrak{N}$ (see (5.17)) can be represented as
Analytic (in $t$) branches of the eigenvalues $\lambda_l(t,\boldsymbol{\theta})$ and analytic branches of the eigenvectors $\varphi_l(t,\boldsymbol{\theta})$ of the operator $A(t,\boldsymbol{\theta})$ admit power series expansions of the form (1.4) and (1.5) with coefficients depending on $\boldsymbol{\theta}$:
The vectors $\omega_1(\boldsymbol{\theta}),\dots,\omega_n(\boldsymbol{\theta})$ form an orthonormal basis in the subspace $\mathfrak{N}$, and the vectors $\zeta_l(\boldsymbol{\theta})=f \omega_l(\boldsymbol{\theta})$, $l=1,\dots,n$, form a basis in $\widehat{\mathfrak{N}}$ (see (6.2)) which is orthonormal with weight: $(\overline{Q}\zeta_l(\boldsymbol{\theta}),\zeta_j(\boldsymbol{\theta}))= \delta_{jl}$, $j, l=1,\dots,n$.
The numbers $\gamma_l(\boldsymbol{\theta})$ and the elements $\omega_l(\boldsymbol{\theta})$ are the eigenvalues and eigenvectors of the spectral germ $S(\boldsymbol{\theta})$. However, it is more convenient to proceed to a generalized spectral problem for $\widehat{S}(\boldsymbol{\theta})$. According to (3.12), the numbers $\gamma_l(\boldsymbol{\theta})$ and the elements $\zeta_l(\boldsymbol{\theta})$ are the eigenvalues and eigenvectors of the following generalized spectral problem:
Now the operators $\widehat{Z}_Q$ and $\widehat{N}_Q$ (defined in § 3.2 in abstract terms) depend on $\boldsymbol{\theta}$. To describe them we introduce the $\Gamma$-periodic solution $\Lambda_Q(\mathbf{x})$ of the problem
Note that the Hermitian matrix-valued function $L_Q(\mathbf{k}):=tL_Q(\boldsymbol{\theta})$, $\mathbf{k} \in \mathbb{R}^d$, is homogeneous of first order. We put $\widehat{N}_Q(\mathbf{k}):= t^3 \widehat{N}_Q(\boldsymbol{\theta})$, $\mathbf{k} \in \mathbb{R}^d$. Then
The matrix-valued function $b(\mathbf{k})^* L_Q(\mathbf{k})b(\mathbf{k})$ is a third-order homogeneous polynomial of $\mathbf{k} \in \mathbb{R}^d$.
Some conditions ensuring that the operator (9.10) is equal to zero were presented in [9].
Proposition 9.1 ([9], § 5). Suppose that at least one of the following assumptions is satisfied:
(a) The operator $\mathcal{A}$ is of the form $\mathcal{A}= f(\mathbf{x})^*\mathbf{D}^*g(\mathbf{x})\mathbf{D}f(\mathbf{x})$, where $g(\mathbf{x})$ is a symmetric matrix with real entries;
(b) Relations (6.20) are satisfied, that is, $g^0=\overline{g}$.
Then $\widehat{N}_Q (\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Proposition 9.2. Suppose that $b(\boldsymbol{\theta})$, $g(\mathbf{x})$, and $Q(\mathbf{x})$ are matrices with real entries. Suppose that in the expansions (9.6) for analytic branches of the eigenvectors of the operator $A(t,\boldsymbol{\theta})$ the ‘embryos’ $\omega_l(\boldsymbol{\theta})$, $l=1,\dots,n$, can be chosen so that the vectors $\zeta_l(\boldsymbol{\theta})=f\omega_l(\boldsymbol{\theta})$ are real. Then the coefficients $\mu_l(\boldsymbol{\theta})$, $l=1,\dots,n$, in (9.5) are equal to zero, that is, $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
In the ‘real’ case under consideration, the germ $\widehat{S}(\boldsymbol{\theta})$ is a symmetric matrix with real entries; $\overline{Q}$ is also a symmetric matrix with real entries. Clearly, in the case of a simple eigenvalue $\gamma_j(\boldsymbol{\theta})$ of the generalized problem (9.7) the eigenvector $\zeta_j(\boldsymbol{\theta})=f\omega_j (\boldsymbol{\theta})$ is defined uniquely up to a phase factor, and we can always choose it to be real. We obtain the following result.
Corollary 9.3. Suppose that the matrices $b(\boldsymbol{\theta})$, $g(\mathbf{x})$, and $Q(\mathbf{x})$ have real entries. Suppose that the spectrum of the problem (9.7) is simple. Then $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
We describe the operators $\widehat{Z}_{2,Q}$, $\widehat{R}_{2,Q}$, and $\widehat{N}_{1,Q}^0$ (defined in abstract terms in § 3.3) for the family $A(t,\boldsymbol{\theta})$. Now these operators depend on the parameter $\boldsymbol{\theta}$. Let $\Lambda_{l,Q}^{(2)}({\mathbf x})$ be the $\Gamma$-periodic solution of the problem
9.4. The multiplicities of eigenvalues of the germ
In the present subsection it is assumed that $n \geqslant 2$. We turn to the notation adopted in § 1.6, keeping track of the multiplicities of eigenvalues of the spectral germ $S(\boldsymbol{\theta})$. The same quantities are eigenvalues of the generalized problem (9.7). In general, the number $p(\boldsymbol{\theta})$ of different eigenvalues $\gamma^{\circ}_1(\boldsymbol{\theta}),\dots, \gamma^{\circ}_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ of the spectral germ $S(\boldsymbol{\theta})$ and their multiplicities $k_1(\boldsymbol{\theta}),\dots, k_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ depend on the parameter $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. For each fixed $\boldsymbol{\theta}$ we denote by $\mathfrak{N}_j(\boldsymbol{\theta})$ the eigenspace of the germ $S(\boldsymbol{\theta})$ corresponding to the eigenvalue $\gamma^{\circ}_j(\boldsymbol{\theta})$. Then $f \mathfrak{N}_j(\boldsymbol{\theta})= \operatorname{Ker}\bigl(\widehat{S}(\boldsymbol{\theta})- \gamma_j^\circ(\boldsymbol{\theta})\overline{Q}\bigr)=: \widehat{\mathfrak N}_{j,Q}(\boldsymbol{\theta})$ is the eigenspace of the problem (9.7) corresponding to the same eigenvalue $\gamma^{\circ}_j(\boldsymbol{\theta})$. We denote by $\mathcal{P}_j(\boldsymbol{\theta})$ the ‘skew’ projection of the space $L_2(\Omega;\mathbb{C}^n)$ onto $\widehat{\mathfrak N}_{j,Q}(\boldsymbol{\theta})$; $\mathcal{P}_j(\boldsymbol{\theta})$ is orthogonal with respect to the inner product with weight $\overline{Q}$. According to (3.15), we have the following invariant representations for the operators $\widehat{N}_{0,Q}(\boldsymbol{\theta})$ and $\widehat{N}_{*,Q}(\boldsymbol{\theta})$:
9.5. The coefficients ${\nu}_l(\boldsymbol{\theta})$
The coefficients ${\nu}_l(\boldsymbol{\theta})$, $l=1,\dots,n$, in the expansions (9.5) are eigenvalues of some problem. We need to describe this problem in the case where ${\mu}_l(\boldsymbol{\theta})=0$, $l=1,\dots,n$, that is, $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$. Applying Proposition 3.3, we obtain the following statement. See also [32], Proposition 11.4.
Proposition 9.4. Suppose that $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$. Let ${\gamma}_1^\circ(\boldsymbol{\theta}),\dots, {\gamma}_{p(\boldsymbol{\theta})}^\circ(\boldsymbol{\theta})$ be the different eigenvalues of the problem (9.7), and let $k_1(\boldsymbol{\theta}),\dots, k_{p(\boldsymbol{\theta})}(\boldsymbol{\theta})$ be their multiplicities. Let $\widehat{P}_{q,Q}(\boldsymbol{\theta})$ be the orthogonal projection of the space $L_2(\Omega;\mathbb{C}^n)$ onto the subspace $\widehat{\mathfrak{N}}_{q,Q}(\boldsymbol{\theta})= \operatorname{Ker}\bigl(\widehat{S}(\boldsymbol{\theta})- {\gamma}_q^\circ(\boldsymbol{\theta})\overline{Q}\bigr)$, $q=1,\dots,p(\boldsymbol{\theta})$. Suppose that the operators $\widehat{Z}_Q(\boldsymbol{\theta})$, $\widehat{N}_Q(\boldsymbol{\theta})$, and $\widehat{N}_{1,Q}^0(\boldsymbol{\theta})$ are defined by (9.9), (9.10), and (9.14), respectively. We introduce the operators $\widehat{\mathcal{N}}^{(q)}_Q(\boldsymbol{\theta})$, $q=1,\dots,p(\boldsymbol{\theta})$: the operator $\widehat{\mathcal{N}}^{(q)}_Q(\boldsymbol{\theta})$ acts on $\widehat{\mathfrak{N}}_{q,Q}(\boldsymbol{\theta})$ and is given by
Denote $i(q,\boldsymbol{\theta})=k_1(\boldsymbol{\theta})+\cdots +k_{q-1}(\boldsymbol{\theta})+1$. Let ${\nu}_l(\boldsymbol{\theta})$, $l=1,\dots,n,$ be the coefficients of $t^4$ from the expansions (9.5), and let $\omega_l(\boldsymbol{\theta})$ be the ‘embryos’ from (9.6). Let ${\zeta}_l(\boldsymbol{\theta})=f {\omega}_l(\boldsymbol{\theta})$, $l=1,\dots,n$. Set $Q_{\widehat{\mathfrak{N}}_{q,Q}(\boldsymbol{\theta})}= \widehat{P}_{q,Q}(\boldsymbol{\theta}) Q\big|_{\widehat{\mathfrak{N}}_{q,Q} (\boldsymbol{\theta})}$. Then
We apply theorems from § 4 to the operator ${A}(t,\boldsymbol{\theta})={\mathcal{A}}(\mathbf{k})$. By Remark 2.10 we can track the dependence of the constants in estimates on the problem data. Note that ${c}_*$, ${\delta}$, and ${t}_0$ do not depend on $\boldsymbol{\theta}$ (see (5.15), (5.24), and (5.26)). According to (5.25), the norm $\|{X}_1(\boldsymbol{\theta})\|$ can be replaced by $\alpha_1^{1/2}\|g\|_{L_{\infty}}^{1/2}\|f\|_{L_{\infty}}$. Therefore, the constants from Theorem 4.1 (as applied to the operator ${\mathcal{A}}(\mathbf{k})$) do not depend on $\boldsymbol{\theta}$. They depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 10.1 ([25]). Let ${\mathcal J}(\mathbf{k},\tau)$ be the operator defined by (10.1). For $\tau \in \mathbb{R}$, $\varepsilon >0$, and $\mathbf k \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_1$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 10.1 is deduced from Theorem 4.1 and relations (7.2)–(7.4). We also take into account the relation
Previously, estimate (10.2) was obtained in [25], Theorem 8.1.
Now we improve the result of Theorem 10.1 under certain additional assumptions. We impose the following condition.
Condition 10.2. Let $\widehat{N}_Q(\boldsymbol{\theta})$ be the operator defined by (9.10). Suppose that $\widehat{N}_Q(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
From Theorem 4.2 we deduce the following result obtained in [30], Theorem 8.2.
Theorem 10.3 ([30]). Let ${\mathcal J}(\mathbf{k},\tau)$ be the operator defined by (10.1). Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_2$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Now we drop Condition 10.2, but assume instead that $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta}$. Assume that $\widehat{N}_Q(\boldsymbol{\theta})= \widehat{N}_{*,Q}(\boldsymbol{\theta}) \ne 0$ for some $\boldsymbol{\theta}$ (otherwise Theorem 10.3 applies). As in § 7.1, in order to apply the ‘abstract’ result of Theorem 4.3, we need to impose some additional conditions. We use the original numbering of the eigenvalues $\gamma_1(\boldsymbol{\theta}),\dots,\gamma_n(\boldsymbol{\theta})$ of the germ $S(\boldsymbol{\theta})$, agreeing to number them in the non-decreasing order:
As already mentioned, the numbers (10.5) are simultaneously the eigenvalues of the generalized spectral problem (9.7). For each $\boldsymbol{\theta}$ we denote by $\mathcal{P}^{(k)}(\boldsymbol{\theta})$ the ‘skew’ projection (orthogonal with weight $\overline{Q}$) of the space $L_2(\Omega;\mathbb{C}^n)$ onto the eigenspace of problem (9.7) corresponding to the eigenvalue $\gamma_k(\boldsymbol{\theta})$. Clearly, for each $\boldsymbol{\theta}$ the operator $\mathcal{P}^{(k)}(\boldsymbol{\theta})$ coincides with one of the projections $\mathcal{P}_j(\boldsymbol{\theta})$ introduced in § 9.4 (but the index $j$ can depend on $\boldsymbol{\theta}$ and changes at points where the multiplicity of the spectrum of the germ changes).
Condition 10.4. $1^\circ$. The operator $\widehat{N}_{0,Q}(\boldsymbol{\theta})$ defined by (9.15) is equal to zero: $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
$2^\circ$. For each pair of indices $(k,r)$, $1 \leqslant k,r \leqslant n$, $k \ne r$, such that $\gamma_k(\boldsymbol{\theta}_0)=\gamma_r(\boldsymbol{\theta}_0)$ for some $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$, we have
for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Obviously, Condition 10.4 is ensured by the following stronger condition.
Condition 10.5. $1^\circ$. The operator $\widehat{N}_{0,Q}(\boldsymbol{\theta})$ defined by (9.15) is equal to zero: $\widehat{N}_{0,Q}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
$2^\circ$. The number $p$ of different eigenvalues of the generalized spectral problem (9.7) does not depend on $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
Remark 10.6. Assumption $2^\circ$ of Condition 10.5 is a fortiori fulfilled if the spectrum of problem (9.7) is simple for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$.
So we assume that Condition 10.4 is satisfied. We are interested in pairs of indices in the set
Since the operator $S(\boldsymbol{\theta})$ depends on $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$ continuously, by the perturbation theory of a discrete spectrum the $\gamma_j(\boldsymbol{\theta})$ are continuous functions on the sphere $\mathbb{S}^{d-1}$. By Condition 10.4($2^\circ$), we have $|\gamma_k(\boldsymbol{\theta})-\gamma_r(\boldsymbol{\theta})| > 0$ for $(k,r) \in \mathcal{K}$ and any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$, so that
Clearly, the number (10.6) is a realization of the value (2.1) chosen independently of $\boldsymbol{\theta}$. Under Condition 10.4, the number $t^{00}$ subject to (2.2) can also be chosen to be independent of $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$. Taking (5.24) and (5.25) into account, we put
where $c^{\circ}$ is defined by (10.6). (The condition $t^{00}\leqslant t_{0}$ is satisfied automatically because $c^{\circ} \leqslant \|S(\boldsymbol{\theta})\| \leqslant \alpha_1\|g\|_{L_{\infty}}\|f\|_{L_\infty}^2$.)
Under Condition 10.4, we deduce the following result from Theorem 4.3 (see [30], Theorem 8.6). We must take into account that now the constants in estimates depend not only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$, but also on $c^\circ$ and $n$; see Remark 2.10.
Theorem 10.7 ([30]). Let ${\mathcal J}(\mathbf{k},\tau)$ be the operator defined by (10.1). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_3$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $c^\circ$ and $n$.
10.2. More accurate approximation in the operator norm on $L_2(\Omega;\mathbb{C}^n)$
Now we obtain more accurate approximation for the sandwiched operator exponential of the operator $A(t,\boldsymbol{\theta})={\mathcal A}(\mathbf k)$ with the help of Theorem 4.4.
The operator (10.12) is bounded, and the operator (10.13) is in the general case defined on $\widetilde{H}^3({\Omega};\mathbb{C}^n)$. From (9.2), (10.4), (10.7), (10.8), and (10.10)–(10.13) it follows that for $\varepsilon>0$, $\tau \in \mathbb{R}$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal{C}}'_4$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Estimates for $|\mathbf{k}| > {t}_0$ are trivial. Taking (7.2) and (10.15) into account we have
By analogy with the proof of estimate (7.27), using the discrete Fourier transform and relations (7.4), (7.6), (9.2), (10.4), and (10.11), it is easy to check that
Comparing estimates (10.16)–(10.18) we arrive at the following result.
Theorem 10.8. Let ${\mathcal G}(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.13). Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_4$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Now, using Theorems 4.5 and 4.6, we improve the result of Theorem 10.8 under certain additional assumptions.
Theorem 10.9. Let ${\mathcal G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.12). Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_5$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Proof. Applying Theorem 4.5 and using Remark 2.10 and relations (7.2), (10.3), and (10.7) we obtain
The constant ${\mathcal{C}}'_5$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
For $|\mathbf{k}| > {t}_0$ we use (7.2) and (10.14):
In combination with (10.20) and (10.21), this yields the required estimate (10.19). $\Box$
In a similar way Theorem 4.6 easily implies the following result; cf. the proof of Theorem 7.11.
Theorem 10.10. Let ${\mathcal G}(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.13). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_6$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^\circ$.
10.3. Approximation of the sandwiched operator exponential in the ‘energy’ norm
Theorem 10.11. Let ${\mathcal G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.12). Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_7$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Proof. Applying Theorem 4.7 and using Remark 2.10 we obtain
The constant ${\mathcal C}_7'$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\| f \|_{L_\infty}$, $\| f^{-1}\|_{L_\infty}$, and $r_0$.
For $|\mathbf{k}| > {t}_0$ estimates are trivial. By (7.2) and (10.26) we have
In a similar way, from Theorems 4.8 and 4.9 we deduce the following two theorems, which improve the result of Theorem 10.11 under certain additional assumptions (cf. the proofs of Theorems 7.13 and 7.14).
Theorem 10.12. Let ${\mathcal G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.12). Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_8$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 10.13. Let ${\mathcal G}_0(\mathbf{k},\varepsilon^{-2}\tau)$ be the operator defined by (10.12). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$, $\varepsilon > 0$, and $\mathbf{k} \in \widetilde{\Omega}$ we have
The constant ${\mathcal C}_9$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^\circ$.
11. Confirmation of the sharpness of the results on approximations for the sandwiched operator $e^{- i \varepsilon^{-2}\tau \mathcal{A}(\mathbf{k})}$
11.1. Sharpness with respect to the smoothing factor
In the statements of the present section we impose one of the following two conditions.
Condition 11.1. Let $\widehat{N}_{0,Q}(\boldsymbol{\theta})$ be the operator defined by (9.15). Then suppose that $\widehat{N}_{0,Q}(\boldsymbol{\theta}_0) \ne 0$ at some point $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$.
Condition 11.2. Let $\widehat{N}_{0,Q}(\boldsymbol{\theta})$ and $\widehat{\mathcal N}^{(q)}_{Q}(\boldsymbol{\theta})$ be the operators defined by (9.15) and (9.16), respectively. Suppose that $\widehat{N}_{0,Q}(\boldsymbol{\theta})\!=\!0$ for any $\boldsymbol{\theta}\! \in\! \mathbb{S}^{d-1}$ and $\widehat{\mathcal N}^{(q)}_{Q}(\boldsymbol{\theta}_0) \!\ne\! 0$ for some $\boldsymbol{\theta}_0 \in \mathbb{S}^{d-1}$ and some $q \in \{1,\dots,p(\boldsymbol{\theta}_0)\}$.
We need the following lemma (see [29], Lemma 11.8, and [32], Lemma 13.3).
Lemma 11.3 ([29], [32]). Let $\delta$ and $t_0$ be defined by (5.24) and (5.26), respectively. Suppose that $F(\mathbf{k})=F(t,\boldsymbol{\theta})$ is the spectral projection of the operator $\mathcal{A}(\mathbf{k})$ for the interval $[0,\delta]$. Then for $|\mathbf{k}| \leqslant t_0$ and $|\mathbf{k}_0| \leqslant t_0$ we have
holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
Proof. Statement $1^\circ$ was proved in [29], Theorem 11.7, on the basis of the abstract result of Theorem 4.10, ($1^\circ$).
Let us prove statement $2^\circ$. We proceed by contradiction. Suppose that for some $\tau \ne 0$ and some $2 \leqslant s < 6$ there exists a constant $\mathcal{C}(\tau)$ such that estimate (11.2) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$. Multiplying the operator under the norm sign in (11.2) by $\widehat{P}$ and using (7.2), we see that
$$
\begin{equation}
f^{-1}\Lambda_Q b(\mathbf{k})\widehat{P}= f^{-1}t\widehat{Z}_Q(\boldsymbol{\theta})\widehat{P}= t {Z}(\boldsymbol{\theta})f^{-1}\widehat{P}.
\end{equation}
\tag{11.6}
$$
Since $f^{-1}\widehat{P}=P f^{-1}\widehat{P}$, the first term on the right-hand side of (11.5) can be written as $f e^{-i \varepsilon^{-2} \tau {\mathcal{A}}(\mathbf{k})} \bigl(P+t{Z}(\boldsymbol{\theta})P\bigr)f^{-1}\widehat{P}$.
Let $|\mathbf k| \leqslant {t}_0$. Then by (1.9) and (1.12),
From Lemma 11.3 (as applied to ${\mathcal{A}} (\mathbf{k})$ and ${\mathcal{A}}^0 (\mathbf{k})$) it follows that, for fixed $\tau$ and $\varepsilon$, the operator ${\mathfrak G}(\mathbf{k},\varepsilon^{-2}\tau)$ is continuous with respect to $\mathbf{k}$ in the ball $|\mathbf{k}| \leqslant {t}_0$. Hence estimate (11.8) is valid for any $\mathbf{k}$ in this ball. In particular, it is true at the point $\mathbf{k}=t\boldsymbol{\theta}_0$ if $t \leqslant {t}_0$. Applying (11.7) again, we see that for some constant $\check{\mathcal{C}}(\tau)>0$ the estimate
holds for any $t \leqslant {t}_0$ and sufficiently small $\varepsilon$.
Estimate (11.9) corresponds to the abstract estimate (4.5). Since $\widehat{N}_{0,Q}(\boldsymbol{\theta}_0)\ne 0$ by Condition 11.1, the assumptions of Theorem 4.10 ($2^\circ$) are satisfied. Applying this theorem, we arrive at a contradiction.
We proceed to the proof of statement $3^\circ$. We prove by contradiction. Suppose that for some $\tau \ne 0$ and $2 \leqslant s < 4$ there exists a constant $\mathcal{C}(\tau)$ such that estimate (11.3) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$. Multiplying the operator under the norm sign in (11.3) by $\widehat{P}$ and taking (7.2) into account, we see that the estimate
for almost all $\mathbf{k}$ in the ball $|\mathbf{k}| \leqslant {t}_0$ and sufficiently small $\varepsilon > 0$. From Lemma 11.3 (as applied to $\mathcal{A}(\mathbf{k})$ and $\mathcal{A}^0(\mathbf{k})$) it follows that for fixed $\tau$ and $\varepsilon$ the operator under the norm sign in (11.13) is continuous with respect to $\mathbf{k}$ in the ball $|\mathbf{k}| \leqslant {t}_0$. Hence estimate (11.13) is valid for all $\mathbf{k}$ in this ball. In particular, it holds for $\mathbf{k}=t\boldsymbol{\theta}_0$ if $t \leqslant {t}_0$. Applying (11.12) once again, for some constant $\check{\mathcal{C}}'(\tau)$ we obtain
for all $t \leqslant {t}_0$ and sufficiently small $\varepsilon$.
Estimate (11.14) corresponds to the abstract estimate (4.6). Since $\widehat{N}_{0,Q}(\boldsymbol{\theta}_0)\ne 0$ by Condition 11.1, the assumptions of Theorem 4.10$(3^\circ$) are satisfied. Applying this theorem, we arrive at a contradiction. $\Box$
In a similar way Theorem 4.11 implies the following result, which confirms the sharpness of Theorems 10.3, 10.7, 10.9, 10.10, 10.12, and 10.13 (on improvements of general results under additional assumptions). Statement $1^\circ$ was obtained in [30], Theorem 8.8.
Theorem 11.5. Suppose that Condition 11.2 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (11.1) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C}(\tau)$, such that estimate (11.2) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$, such that estimate (11.3) holds for almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
11.2. The sharpness of results with respect to time
In the present subsection we confirm the sharpness of the results of § 10 with respect to the dependence of estimates on $\tau$ (for large $|\tau|$).
By analogy with the proof of Theorem 11.4, from Theorem 4.12 we deduce the following statement, which confirms that Theorems 10.1, 10.8, and 10.11 are sharp. Statement $1^\circ$ was obtained in [30], Theorem 8.9.
Theorem 11.6. Suppose that Condition 11.1 is satisfied.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau|=0$ and estimate (11.1) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$, and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/\tau^2=0$ and estimate (11.2) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/|\tau|=0$ and estimate (11.3) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
In a similar way Theorem 4.13 implies the following result, which shows that Theorems 10.3, 10.7, 10.9, 10.10, 10.12, and 10.13 are sharp. Statement $1^\circ$ was obtained in [30], Theorem 8.10.
Theorem 11.7. Suppose that Condition 11.2 is satisfied. Then the following hold.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau)/|\tau|^{1/2}=0$ and estimate (11.1) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau| = 0$ and estimate (11.2) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau)/|\tau|^{1/2} =0$ and estimate (11.3) holds for all $\tau \in \mathbb{R}$, almost all $\mathbf{k} \in \widetilde{\Omega}$ and sufficiently small $\varepsilon > 0$.
12. Approximation of the operator $e^{-i \varepsilon^{-2} \tau \widehat{\mathcal{A}}}$
12.1. Approximation of the operator $e^{-i \varepsilon^{-2} \tau \widehat{\mathcal{A}}}$ in the principal order
In $L_2(\mathbb{R}^d;\mathbb{C}^n)$ consider the operator
Therefore, Theorems 7.1, 7.3, and 7.8 imply the following statements directly. For brevity, below we combine formulations (on improvements of results), so it is convenient to start a new numbering of constants.
Theorem 12.1 ([25]). For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_1$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Previously, estimate (12.4) was obtained in [25], Theorem 9.1.
Theorem 12.2 ([30]). Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 7.2 the constant $\widehat{\mathrm{C}}_2$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 7.4 this constant depends on the same parameters and also on $n$ and $\widehat{c}^{\circ}$.
Estimate (12.5) was obtained in [30], Theorems 9.2 and 9.3.
12.2. More accurate approximation
We need the operator $\Pi=\mathcal{U}^{-1}[\widehat{P}]\mathcal{U}$ acting on $L_2(\mathbb{R}^d;\mathbb{C}^n)$. Here $[\widehat{P}]$ is the orthogonal projection in $\mathcal{H}\!=\! \int_{\widetilde{\Omega}} \!\oplus L_2 (\Omega; \mathbb{C}^n) \, d \mathbf{k}$, acting on fibres of the direct integral as the operator $\widehat{P}$ of averaging over the cell. In [8], formula (6.8), it was shown that $\Pi$ is given by
where $\widehat{\mathbf{u}}(\boldsymbol{\xi})$ is the Fourier image of the function $\mathbf{u} (\mathbf{x})$. Thus, $\Pi$ is the pseudodifferential operator in $L_2(\mathbb{R}^d;\mathbb{C}^n)$, whose symbol is the characteristic function $\chi_{\widetilde{\Omega}}(\boldsymbol{\xi})$ of the set $\widetilde{\Omega}$.
The operator (12.6) is bounded, and the operator (12.7) is in the general case defined on $H^3(\mathbb{R}^d;\mathbb{C}^n)$. Below we will see that under Condition 7.4 the operator (12.7) is defined on $H^1(\mathbb{R}^d;\mathbb{C}^n)$ (this follows from representation (12.9) and Proposition 12.6). The operators (12.6) and (12.7) can be represented as
Let $\widehat{G}_0(\mathbf k,\varepsilon^{-2}\tau)$ and $\widehat{G}(\mathbf k,\varepsilon^{-2}\tau)$ be the operators defined by (7.14) and (7.15). The operators (12.6) and (12.7) expand in direct integrals:
The constant $\widehat{\mathrm{C}}_3$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 12.4. Let $\widehat{G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (12.6). Suppose that Condition 7.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_4$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 12.5. Let $\widehat{G}(\varepsilon^{-2} \tau)$ be the operator given by (12.7). Suppose that Condition 7.4 (or more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_5$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^{\circ}$.
For the purposes of interpolation, in Chapter 3 we need the following two statements.
Proposition 12.6. Suppose that Condition 7.4 is satisfied. Then the operator $\widehat{G}^{(3)}(\varepsilon^{-2}\tau)$ defined by (12.11) can be represented as the pseudodifferential operator with symbol
Here ${\boldsymbol{\xi}}=|{\boldsymbol{\xi}}|\widehat{\boldsymbol{\xi}} \in \mathbb{R}^d$, $\widehat{\boldsymbol{\xi}} \in \mathbb{S}^{d-1}$, the numbers $\widehat\gamma_l^\circ(\widehat{\boldsymbol{\xi}})$, $l=1,\dots,p(\widehat{\boldsymbol{\xi}})$, are the different eigenvalues of the matrix $\widehat S(\widehat{\boldsymbol{\xi}})= b(\widehat{\boldsymbol{\xi}})^* g^0 b(\widehat{\boldsymbol{\xi}})$, and $\widehat{P}_l(\widehat{\boldsymbol{\xi}})$ is the orthogonal projection of $\mathbb{C}^n$ onto the eigenspace of the operator $\widehat S(\widehat{\boldsymbol{\xi}})$ corresponding to the eigenvalue $\widehat\gamma_l^\circ(\widehat{\boldsymbol{\xi}})$. We have
Under Condition 7.4 we have $\widehat{N}(\widehat{\boldsymbol{\xi}})= \widehat{N}_*(\widehat{\boldsymbol{\xi}})$. Therefore, according to (6.30), the matrix $\widehat{N}(\widehat{\boldsymbol{\xi}})= b(\widehat{\boldsymbol{\xi}})^* L(\widehat{\boldsymbol{\xi}}) b(\widehat{\boldsymbol{\xi}})$ can be written as
Since $\widehat{P}_l (\widehat{\boldsymbol{\xi}})$ is the orthogonal projection of $\mathbb{C}^n$ onto the eigenspace of the operator $\widehat S(\widehat{\boldsymbol{\xi}})$ corresponding to the eigenvalue $\widehat{\gamma}_l^\circ(\widehat{\boldsymbol{\xi}})$, it follows that
Calculating the integrals on the right, we arrive at representation (12.14). Estimate (12.15) follows from (12.14) and the estimates $|b(\widehat{\boldsymbol{\xi}})^* L(\widehat{\boldsymbol{\xi}}) b(\widehat{\boldsymbol{\xi}})| \leqslant C_{\widehat{N}}$ and $|\widehat\gamma_j^\circ(\widehat{\boldsymbol{\xi}})- \widehat\gamma_l^\circ(\widehat{\boldsymbol{\xi}})| \geqslant \widehat{c}^\circ$, $j \ne l$. $\Box$
Proposition 12.7. Let $\widehat{G}_0(\varepsilon^{-2}\tau)$ and $\widehat{G}( \varepsilon^{-2}\tau)$ be the operators defined by (12.6) and (12.7). Then the following hold.
$1^\circ$. For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants $\widehat{\mathrm C}_6^\circ$ and $\widehat{\mathrm C}_6$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
$2^\circ$. Suppose that Condition 7.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm C}_7$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
$3^\circ$. Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants $\widehat{\mathrm C}_8^\circ$ and $\widehat{\mathrm C}_8$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$, and also on $n$ and $\widehat{c}^\circ$.
Proof. $1^\circ$. We use representations (12.8) and (12.9). By Theorem 12.1 the operator $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal{A}}}- e^{-i\varepsilon^{-2}\tau\widehat{\mathcal{A}}^{\,0}}$ satisfies estimate (12.4).
Note that the operator (12.10) expands in a direct integral of the operators (7.18). Combining this with (7.2), (7.20), and the relation $\widehat{G}^{(2)}(\mathbf{k},\varepsilon^{-2}\tau)= \widehat{G}^{(2)}(\mathbf{k},\varepsilon^{-2}\tau)\widehat{P}$ we obtain
$2^\circ$. We use representation (12.8). Under Condition 7.2, by Theorem 12.2 the operator $e^{-i\varepsilon^{-2} \tau \widehat{\mathcal{A}}} -e^{-i \varepsilon^{-2} \tau \widehat{\mathcal{A}}^{\,0}}$ satisfies estimate (12.5).
$3^\circ$. We use representations (12.8) and (12.9). Under Condition 7.4, by Theorem 12.2 the operator $e^{-i\varepsilon^{-2} \tau \widehat{\mathcal{A}}} -e^{-i\varepsilon^{-2} \tau \widehat{\mathcal{A}}^{\,0}}$ satisfies estimate (12.5). Estimate (12.22) for the operator $\widehat{G}^{(2)}(\varepsilon^{-2}\tau)\mathcal{R}(\varepsilon)$ remains true. Relations (12.5), (12.8), and (12.22) imply (12.19).
Now we apply Proposition 12.6 under Condition 7.4, which yields
Therefore, Theorems 7.12–7.14 imply the following statements.
Theorem 12.8. Let $\widehat{G}_0(\varepsilon^{-2} \tau)$ be the operator defined by (12.6). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_9$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 12.9. Let $\widehat{G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (12.6). Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 7.2, the constant $\widehat{\mathrm{C}}_{10}$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 7.4, this constant depends on the same parameters and also on $n$ and $\widehat{c}^{\circ}$.
For the purposes of interpolation, in Chapter 3 we also need the following statement.
Proposition 12.10. Let $\widehat{G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (12.6). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants $\widehat{\mathrm C}_{11}^\circ$ and $\widehat{\mathrm C}_{11}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and $r_1$.
Proof. We use representation (12.8). Obviously, we have
holds for all sufficiently small $\varepsilon > 0$.
Proof. We prove statement $1^\circ$ by contradiction. Suppose that for some $\tau \ne 0$ and $0 \leqslant s<3$ there exists a constant $\mathcal{C}(\tau) > 0$ such that estimate (12.30) holds for all sufficiently small $\varepsilon > 0$. By (12.3) this means that estimate (8.1) is valid for almost all $\mathbf{k} \in \widetilde{\Omega}$ and all sufficiently small $\varepsilon$. But this contradicts statement $1^\circ$ of Theorem 8.4.
In a similar way statements $2^\circ$ and $3^\circ$ are deduced from statements $2^\circ$ and $3^\circ$ of Theorem 8.4, respectively. $\Box$
Previously, statement $1^\circ$ was obtained in [29], Theorem 12.4.
The sharpness of the improved results (Theorems 12.2, 12.4, 12.5, and 12.9) follows from Theorem 8.5.
Theorem 12.12. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (12.30) holds for all sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C} (\tau)$ such that estimate (12.31) holds for all sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (12.32) holds for all sufficiently small $\varepsilon > 0$.
Statement $1^\circ$ was obtained in [30], Theorem 9.5.
We proceed to sharpness with respect to the dependence of estimates on the parameter $\tau$. Theorem 8.6 implies the sharpness of the general results (Theorems 12.1, 12.3, and 12.8).
Theorem 12.13. Suppose that Condition 8.1 is satisfied.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau| = 0$ and estimate (12.30) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /\tau^2 = 0$ and estimate (12.31) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (12.32) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 9.6.
The sharpness of the improved results (Theorems 12.2, 12.4, 12.5, and 12.9) follows from Theorem 8.7.
Theorem 12.14. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau|^{1/2} =0$ and estimate (12.30) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (12.31) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau|^{1/2} =0$ and estimate (12.32) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was checked in [30], Theorem 9.7.
13. Approximation for the sandwiched exponential $e^{-i\varepsilon^{-2}\tau{\mathcal A}}$
13.1. Approximation for the sandwiched operator $e^{-i \varepsilon^{-2} \tau {\mathcal A} }$ in the principal order
In $L_2 (\mathbb{R}^d; \mathbb{C}^n)$ consider the operator (5.10). Let $f_0$ be the matrix (9.1), and let $\mathcal{A}^0$ be the operator (9.3). Denote
Below we also need the notation ${\mathcal J}(\mathbf{k},\tau)$ introduced in (10.1). Combining decompositions of the form (5.21) for ${\mathcal{A}}$ and ${\mathcal{A}}^0$ and (12.2) we obtain
The constant ${\mathrm{C}}_1$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 13.2 ([30]). Let ${\mathcal J}(\tau)$ be the operator defined by (13.1). Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 10.2, the constant ${\mathrm{C}}_2$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 10.4, this constant depends on the same parameters and also on $n$ and ${c}^{\circ}$.
Previously, Theorem 13.1 was obtained in [25], Theorem 10.1, and Theorem 13.2 was proved in [30], Theorems 9.9 and 9.10.
The operator (13.2) is bounded, and the operator (13.3) is in the general case defined on $H^3(\mathbb{R}^d;\mathbb{C}^n)$. Under Condition 10.4 the operator (13.3) is defined on $H^1(\mathbb{R}^d;\mathbb{C}^n)$ (this follows from representation (13.4) and Proposition 13.6 stated below).
Let ${\mathcal G}_0(\mathbf k,\varepsilon^{-2} \tau)$ and ${\mathcal G}(\mathbf k, \varepsilon^{-2} \tau)$ be the operators given by (10.12) and (10.13), respectively. Similarly to (12.12) and (12.13), we have
These relations, in combination with Theorems 10.8, 10.9, and 10.10, imply the following statements directly.
Theorem 13.3. Let ${\mathcal G}(\varepsilon^{-2} \tau)$ be the operator defined by (13.3). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_3$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 13.4. Let ${\mathcal G}_0(\varepsilon^{-2} \tau)$ be the operator defined by (13.2). Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_4$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 13.5. Let ${\mathcal G}(\varepsilon^{-2} \tau)$ be the operator defined by (13.3). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_5$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^{\circ}$.
For the purposes of interpolation, in Chapter 3 we need the following statements. The first of them can easily be checked similarly to the proof of Proposition 12.6.
Proposition 13.6. Suppose that Condition 10.4 is satisfied. Then the operator ${\mathcal G}^{(3)}(\varepsilon^{-2}\tau)$ defined by (13.5) can be represented as the pseudodifferential operator with symbol
Here ${\boldsymbol{\xi}}=|{\boldsymbol{\xi}}|\widehat{\boldsymbol{\xi}} \in \mathbb{R}^d$, $\widehat{\boldsymbol{\xi}} \in \mathbb{S}^{d-1}$, the numbers $\gamma_l^\circ(\widehat{\boldsymbol{\xi}})$, $l= 1,\dots, p(\widehat{\boldsymbol{\xi}})$, are the different eigenvalues of the generalized problem $\widehat S(\widehat{\boldsymbol{\xi}}) {\mathbf c}= \gamma \overline{Q}{\mathbf c}$, ${\mathbf c} \in \mathbb{C}^n$, and ${\mathcal P}_l (\widehat{\boldsymbol{\xi}})$ is the orthogonal projection with weight $\overline{Q}$ of $\mathbb{C}^n$ onto the corresponding eigenspace. We also have
The second statement is proved by analogy with the proof of Proposition 12.7, by using Theorems 13.1 and 13.2 and Proposition 13.6.
Proposition 13.7. Let ${\mathcal G}_0( \varepsilon^{-2}\tau)$ and ${\mathcal G}(\varepsilon^{-2}\tau)$ be the operators defined by (13.2) and (13.3). Then the following hold.
$1^\circ$. For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm C}_6^\circ$ and ${\mathrm C}_6$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
$2^\circ$. Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm C}_7$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
$3^\circ$. Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm C}_8^\circ$ and ${\mathrm C}_8$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^\circ$.
where the operator ${\mathcal G}_0(\mathbf{k},\varepsilon^{-2} \tau)$ is defined by (10.12). Then Theorems 10.11–10.13 imply the following two statements.
Theorem 13.8. Let ${\mathcal G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (13.2). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_9$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 13.9. Let ${\mathcal G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (13.2). Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 10.2 the constant ${\mathrm{C}}_{10}$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 10.4 this constant depends on the same parameters and also on $n$ and ${c}^{\circ}$.
For the purposes of interpolation, in Chapter 3 we also need the following statement, which can easily be checked similarly to the proof of Proposition 12.10.
Proposition 13.10. Let ${\mathcal G}_0(\varepsilon^{-2}\tau)$ be the operator defined by (13.2). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm C}_{11}^\circ$ and ${\mathrm C}_{11}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and $r_1$.
13.4. The sharpness of the results of §§ 13.1–13.3
Theorems of § 11 imply that the results of §§ 13.1–13.3 are sharp. We start with sharpness with respect to the smoothing factor.
The sharpness of the general results (Theorems 13.1, 13.3, and 13.8) follows from Theorem 11.4.
Theorem 13.11. Suppose that Condition 11.1 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s<3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that the inequality
Previously, statement $1^\circ$ was proved in [29], Theorem 12.8.
The sharpness of the improved results (Theorems 13.2, 13.4, 13.5, and 13.9) follows from Theorem 11.5.
Theorem 13.12. Suppose that Condition 11.2 is satisfied. Then the following hold.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (13.7) holds for sufficiently small $\varepsilon>0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C} (\tau)$ such that estimate (13.8) holds for sufficiently small $\varepsilon>0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (13.9) holds for sufficiently small $\varepsilon>0$.
Statement $1^\circ$ was obtained in [30], Theorem 9.12.
We proceed to sharpness with respect to the dependence of estimates on the parameter $\tau$. The sharpness of the general results (Theorems 13.1, 13.3, and 13.8) follows from Theorem 11.6.
Theorem 13.13. Suppose that Condition 11.1 is satisfied. Then the following hold.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau| = 0$ and estimate (13.7) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /\tau^2 = 0$ and estimate (13.8) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty}\mathcal{C}(\tau) / |\tau| = 0$ and estimate (13.9) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was proved in [30], Theorem 9.13.
Finally, the improved results (Theorems 13.2, 13.4, 13.5, and 13.9) are also sharp, which follows from Theorem 11.7.
Theorem 13.14. Suppose that Condition 11.2 is satisfied.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau|^{1/2} =0$ and estimate (13.7) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (13.8) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau|^{1/2} =0$ and estimate (13.9) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was proved in [30], Theorem 9.14.
Chapter 3. Homogenization for Schrödinger-type equations
14. Approximation for the operator $e^{-i\tau\widehat{\mathcal{A}}_\varepsilon}$
14.1. The operators $\widehat{\mathcal{A}}_\varepsilon$ and $\mathcal{A}_\varepsilon$. The statement of the problem
If $\psi(\mathbf{x})$ is a measurable $\Gamma$-periodic function in $\mathbb{R}^d$, then we set $\psi^{\varepsilon}(\mathbf{x}):=\psi(\varepsilon^{-1}\mathbf{x})$, $\varepsilon > 0$. Our main objects are the operators $\widehat{\mathcal{A}}_\varepsilon$ and $\mathcal{A}_\varepsilon$ acting in $L_2(\mathbb{R}^d;\mathbb{C}^n)$ and given formally by
They are defined rigorously in terms of the corresponding quadratic forms (cf. § 5.3). The coefficients of the operators (14.1) and (14.2) oscillate rapidly as $\varepsilon \to 0$.
Our goal is to obtain approximations of the operators $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ and $f^\varepsilon e^{-i\tau\mathcal{A}_\varepsilon}(f^\varepsilon)^{-1}$ for small $\varepsilon$ and to apply the results to homogenization of the solutions of the Cauchy problem for Schrödinger-type equations.
14.2. The scaling transformation
Let $T_{\varepsilon}$ be a unitary scaling transformation in $L_2(\mathbb{R}^d;\mathbb{C}^n)$:
If $\psi(\mathbf{x})$ is a $\Gamma$-periodic measurable function, then, under the scaling transformation, the operator $[\psi^{\varepsilon}]$ of multiplication by the function $\psi^{\varepsilon}(\mathbf{x})$ turns to the operator $[\psi]$ of multiplication by $\psi(\mathbf{x})$: $[\psi^{\varepsilon}]=T_\varepsilon^*[\psi]T_\varepsilon$. We have $\mathcal{A}_\varepsilon = \varepsilon^{-2}T_{\varepsilon}^*\mathcal{A}T_{\varepsilon}$. Hence
14.3. Approximation for the operator $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ in the principal order
Applying (14.3) to the operators $\widehat{\mathcal{A}}_\varepsilon$ and $\widehat{\mathcal{A}}^{\,0}$, and also using (14.4), for $\tau \in \mathbb{R}$ and $\varepsilon>0$ we obtain
Note that the operator $(\mathcal{H}_0+I)^{s/2}$ is an isometric isomorphism of the Sobolev space $H^s(\mathbb{R}^d; \mathbb{C}^n)$ onto $L_2(\mathbb{R}^d;\mathbb{C}^n)$. Hence, using Theorems 12.1 and 12.2 and relation (14.5), we obtain the following two theorems directly.
Theorem 14.1 ([25]). Let $\widehat{\mathcal{A}}_{\varepsilon}$ be the operator (14.1) and let $\widehat{\mathcal{A}}^{\,0}$ be the effective operator (6.17). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_1$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 14.2 ([30]). Suppose that the assumptions of Theorem 14.1 are satisfied. Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 7.2 the constant $\widehat{\mathrm{C}}_2$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 7.4 this constant depends on the same parameters and also on $n$ and $\widehat{c}^\circ$.
Previously, Theorem 14.1 was proved in [25], Theorem 12.1, and Theorem 14.2 was obtained in [30], Theorems 10.2 and 10.3.
By interpolation, Theorems 14.1 and 14.2 imply the following statements.
Corollary 14.3. Under the assumptions of Theorem 14.1 we have
Here $\widehat{\mathfrak{C}}_2(s)=2^{1-s/2}\widehat{\mathrm{C}}_2^{s/2}$.
Proof. Interpolating between (14.9) and (14.7), we arrive at estimate (14.10). $\Box$
14.4. More accurate approximation
Let $\Pi_\varepsilon:=T_\varepsilon^* \Pi T_\varepsilon$. Then $\Pi_\varepsilon$ is the pseudodifferential operator in $L_2(\mathbb{R}^d; \mathbb{C}^n)$ with symbol $\chi_{\widetilde{\Omega}/\varepsilon} ({\boldsymbol \xi})$:
The operator (14.11) is bounded, and the operator (14.12) is in the general case defined on the space $H^3(\mathbb{R}^d;\mathbb{C}^n)$. Under Condition 7.4 the operator (14.12) is defined on $H^1(\mathbb{R}^d;\mathbb{C}^n)$; cf. § 12.2.
Let $\widehat{G}_{0}(\varepsilon^{-2}\tau)$ and $\widehat{G}(\varepsilon^{-2}\tau)$ be the operators defined by (12.6) and (12.7), respectively. Applying the scaling transformation we obtain
The constant $\widehat{\mathrm{C}}_3$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 14.6. Let $\widehat{G}_{0,\varepsilon}(\tau)$ be the operator defined by (14.11). Suppose that Condition 7.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_4$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 14.7. Let $\widehat{G}_\varepsilon(\tau)$ be the operator defined by (14.12). Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm{C}}_5$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and $\widehat{c}^{\circ}$.
In a similar way, from Proposition 12.7 and relations (14.13) and (14.14) we deduce the following statement.
Proposition 14.8. Let $\widehat{G}_{0,\varepsilon}(\tau)$ and $\widehat{G}_\varepsilon(\tau)$ be the operators defined by (14.11) and (14.12). Then the following hold.
$1^\circ$. For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants $\widehat{\mathrm C}_6^\circ$ and $\widehat{\mathrm C}_6$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
$2^\circ$. Suppose that Condition 7.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant $\widehat{\mathrm C}_7$ depends only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
$3^\circ$. Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants $\widehat{\mathrm C}_8^\circ$ and $\widehat{\mathrm C}_8$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$, and also on $n$ and $\widehat{c}^\circ$.
By interpolation, Theorems 14.5–14.7 and Proposition 14.8 imply the following results.
Corollary 14.9. Under the assumptions of Theorem 14.5, we have
We obtain approximation for the operator $e^{-i\tau\widehat{\mathcal{A}}_\varepsilon} \bigl(I+\varepsilon\Lambda^\varepsilon b({\mathbf D})\Pi_\varepsilon\bigr)$ in the $(H^s \to H^1)$-norm (‘energy’ norm) and also approximation for the operator $g^\varepsilon b({\mathbf D}) e^{-i \tau \widehat{\mathcal{A}}_\varepsilon} \bigl(I+\varepsilon\Lambda^\varepsilon b({\mathbf D})\Pi_\varepsilon\bigr)$ (corresponding to the ‘flux’) in the $(H^s \to L_2)$-norm. We put
Using this relation and Theorems 12.8 and 12.9 we deduce the following results.
Theorem 14.12. Let $\widehat{G}_{0,\varepsilon}(\tau)$ and $\widehat{\Xi}_\varepsilon(\tau)$ be the operators (14.11) and (14.26). Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constants $\widehat{\mathrm C}_{12}$ and $\widehat{\mathrm C}_{13}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$.
Proof. Using (14.27) and taking into account that the operator $T_\varepsilon$ is unitary, from Theorem 12.8 we obtain the estimate
Theorem 14.13. Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Let $\widehat{G}_{0,\varepsilon}(\tau)$ be the operator (14.11), and let $\widehat{\Xi}_\varepsilon(\tau)$ be the operator (14.26). Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
Under Condition 7.2 the constants $\widehat{\mathrm C}_{14}$ and $\widehat{\mathrm C}_{15}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 7.4 they depend on the same parameters and also on $n$ and $\widehat{c}^\circ$.
Proof. From (14.27) and Theorem 12.9 it follows that
Proposition 14.14. Let $\widehat{G}_{0,\varepsilon}(\tau)$ be the operator (14.11), and let $\widehat{\Xi}_\varepsilon(\tau)$ be the operator (14.26). Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constants $\widehat{\mathrm C}_{16}$ and $\widehat{\mathrm C}_{17}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and $r_1$.
Remark 14.17. (i) In the general case, that is, under the assumptions of Theorems 14.1, 14.5, and 14.12, we can consider large values of time $\tau=O(\varepsilon^{-\alpha})$, $0< \alpha < 1$, and obtain the qualified estimates
(ii) In the case of improvements of general results, that is, under the assumptions of Theorems 14.2, 14.6, 14.7, and 14.13, we can consider $\tau=O(\varepsilon^{-\alpha})$, $0< \alpha < 2$, and obtain the qualified estimates
The results of §§ 14.4 and 14.5 present an approximation not to the exponential $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon}$ itself, but to the composition $e^{-i \tau \widehat{\mathcal{A}}_\varepsilon} (I+\varepsilon \Lambda^\varepsilon b({\mathbf D}) \Pi_\varepsilon)$. If we could approximate the ‘problematic term’ $e^{-i\tau \widehat{\mathcal{A}}_\varepsilon}\varepsilon \Lambda^\varepsilon b({\mathbf D})\Pi_\varepsilon$ with the required accuracy, then this would lead to approximations for the exponential $e^{-i\tau\widehat{\mathcal{A}}_\varepsilon}$. However, it is impossible to approximate this operator in the same terms as previously (that is, in terms of the spectral characteristics of the operator $\widehat{\mathcal{A}}$ at the edge of the spectrum). Indeed, after the scaling transformation and the direct integral decomposition, the ‘problematic term’ transforms into the operator $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal{A}}(\mathbf k)}[\Lambda] b(\mathbf k)\widehat{P}$ acting on $L_2(\Omega;\mathbb{C}^n)$. Since
within the margin of error $\widehat{P}^\perp$ can be replaced by $\widehat{F}(\mathbf k)^\perp$, and we obtain the ‘new problematic term’ $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal{A}}(\mathbf k)} \widehat{F}(\mathbf k)^\perp[\Lambda]b(\mathbf k)\widehat{P}$. Clearly, the operator $e^{-i\varepsilon^{-2}\tau\widehat{\mathcal{A}} (\mathbf k)} \widehat{F}(\mathbf k)^\perp$ cannot be approximated in ‘threshold’ terms, since $\widehat{F}(\mathbf k)^\perp$ is the spectral projection of the operator $\widehat{\mathcal{A}}(\mathbf k)$ corresponding to the interval $[3\delta,\infty)$.
14.7. The sharpness of the results of §§ 14.3–14.5
Applying theorems from § 12.4 we verify that the results of §§ 14.3–14.5 are sharp. First, we discuss the sharpness with respect to the type of operator norm.
Let us check that the general results (Theorems 14.1, 14.5, and 14.12) are sharp.
Theorem 14.18. Suppose that Condition 8.1 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that the inequality
holds for all sufficiently small $\varepsilon > 0$.
Proof. We check statement $1^\circ$. Suppose that for some $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$ estimate (14.46) is fulfilled for sufficiently small $\varepsilon$. Applying the scaling transformation (see (14.5)) we see that estimate (12.30) also holds. But this contradicts statement $1^\circ$ of Theorem 12.11.
In a similar way statement $2^\circ$ is deduced from statement $2^\circ$ of Theorem 12.11.
Let us check statement $3^\circ$. Suppose that for some $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$ estimate (14.48) holds for sufficiently small $\varepsilon$. Then
is also fulfilled for sufficiently small $\varepsilon$ (for some constant $\widetilde{\mathcal{C}}(\tau)>0$). Applying the scaling transformation (see (14.27)), we see that estimate (12.32) also holds. But this contradicts statement $3^\circ$ of Theorem 12.11. $\Box$
Previously, statement $1^\circ$ was obtained in [29], Theorem 13.6.
In a similar way, using the scaling transformation, Theorem 12.12 implies that the improved results (Theorems 14.2, 14.6, 14.7, and 14.13) are sharp.
Theorem 14.19. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (14.46) holds for sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (14.47) holds for sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (14.48) holds for sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.5.
We proceed to sharpness with respect to the dependence of estimates on the parameter $\tau$. Using the scaling transformation, from Theorem 12.13 we deduce the following result, which confirms that the general results (Theorems 14.1, 14.5, and 14.12) are sharp.
Theorem 14.20. Suppose that Condition 8.1 is satisfied.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (14.46) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / \tau^2 = 0$ and estimate (14.47) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (14.48) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.6.
Finally, using the scaling transformation, from Theorem 12.14 we deduce the following result, which demonstrates that the improved results (Theorems 14.2, 14.6, 14.7, and 14.13) are sharp.
Theorem 14.21. Suppose that Condition 8.2 is satisfied.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau|^{1/2}=0$ and estimate (14.46) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (14.47) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau|^{1/2}=0$ and estimate (14.48) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.7.
14.8. On the possibility to remove the smoothing operator $\Pi_\varepsilon$ from approximations
Now, we consider the question of whether it is possibile to remove the operator $\Pi_\varepsilon$ from approximations (that is, to replace $\Pi_\varepsilon$ by the identity operator while preserving the order of errors) in the results of §§ 14.4 and 14.5.
Lemma 14.22. Let $s\geqslant 1$. Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constant ${\mathrm C}(s)$ depends on $\alpha_1$, $r_0$, and $s$.
Proof. Writing the norm on the left-hand side of (14.49) in the Fourier representation and recalling that the symbol of the operator $\Pi$ is $\chi_{\widetilde{\Omega}}({\boldsymbol \xi})$, we obtain
From Corollary 14.9 and Lemma 14.22 we deduce the following statement. Below $[\Lambda]$ denotes the operator of multiplication by the $\Gamma$-periodic solution of problem (6.8).
Theorem 14.23. Let $\widehat{G}'_\varepsilon(\tau)$ be the operator defined by (14.51). Let $3 \leqslant s \leqslant 6$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant $\widehat{\mathfrak C}'_3(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
Proof. From (14.52), using the scaling transformation we obtain
In the last transition we used estimate (14.49). In combination with Corollary 14.9 and the restriction $0< \varepsilon \leqslant 1$ this implies the required estimate (14.53). $\Box$
In a similar way Corollaries 14.10 and 14.11 and Lemma 14.22 imply the following two results.
Theorem 14.24. Let $\widehat{G}'_{0,\varepsilon}( \tau)$ be the operator defined by (14.50). Suppose that Condition 7.2 is satisfied. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant $\widehat{\mathfrak C}'_4(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
Theorem 14.25. Let $\widehat{G}'_\varepsilon(\tau)$ be the operator defined by (14.51). Suppose that Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant $\widehat{\mathfrak C}'_5(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, $n$, $\widehat{c}^{\circ}$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
Now, we consider the question about the possible removal of the operator $\Pi_\varepsilon$ from approximations of the operator exponential in the ‘energy’ norm. We put
From Corollary 14.15 and Lemma 14.22 we deduce the following statement.
Theorem 14.26. Let $\widehat{G}'_{0,\varepsilon}(\tau)$ and $\widehat{\Xi}_\varepsilon'(\tau)$ be the operators defined by (14.50) and (14.56), respectively. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constants $\widehat{\mathfrak C}'_6(s)$ and $\widehat{\mathfrak C}'_7(s)$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$.
Remark 14.27. Formally, the statement of Theorem 14.26 is valid for $1\leqslant s \leqslant 4$, but for $1\leqslant s <2$ the condition on $\Lambda$ can only be fulfilled in the case where $\Lambda=0$.
Proof of Theorem 14.26. By analogy with (14.54), we have
The constant $\widehat{\mathfrak C}_{6}''(s)$ depends on $\alpha_1, \|g\|_{L_\infty}, r_0, s$, and on the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$. Relations (14.42) and (14.61) imply (14.57).
Consider the second term on the right-hand side of (14.62). Since $\widetilde{g}=g(b({\mathbf D})\Lambda+ {\mathbf 1})$ and the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$, according to [111], § 1.3.2, Lemma 1, the operator $ [\widetilde{g}]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. The norm $\|\kern0.2pt[\widetilde{g}\,]\kern0.2pt\|_{H^{s-1}\to L_2}$ is controlled in terms of $\|\kern0.2pt[\Lambda]\kern0.2pt\|_{H^{s-1}\to H^1}$ and $\|g\|_{L_\infty}$. Using the scaling transformation and Lemma 14.22, we obtain
where the constant $\widehat{\mathfrak C}_{7}''(s)$ depends on $\alpha_1$, $\|g\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|\kern0.2pt[\Lambda]\kern0.2pt\|_{H^{s-1} \to H^1}$. In combination with (14.43) this yields (14.58). The proof is completed.
In a similar way Corollary 14.16 and Lemma 14.22 imply the following result.
Theorem 14.28. Let $\widehat{G}'_{0,\varepsilon}(\tau)$ and $\widehat{\Xi}_\varepsilon'(\tau)$ be the operators defined by (14.50) and (14.56), respectively. Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Let $2 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
Under Condition 7.2 the constants $\widehat{\mathfrak C}'_{8}(s)$ and $\widehat{\mathfrak C}'_9(s)$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$. Under Condition 7.4 these constants depend on the same parameters and also on $n$ and $\widehat{c}^{\circ}$.
Let us discuss some cases where one of the conditions on the operator $[\Lambda]$ (from Theorems 14.23–14.26 and 14.28) is fulfilled. We need the following auxiliary facts.
Proposition 14.29. Let $l \geqslant 0$. Suppose that $\Upsilon$ is a $\Gamma$-periodic function in $\mathbb{R}^d$ such that
$1^\circ$. If $d\leqslant 2s$, then the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$, and the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$ is controlled in terms of $d$, $\alpha_0$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and the parameters of the lattice $\Gamma$.
$2^\circ$. If $d > 2s$, then for the continuity of the operator $[\Lambda]$ from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$ it suffices that $\Lambda$ belongs to $L_{d/(s-1)}(\Omega)$. The norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$ is controlled in terms of $d$, the parameters of the lattice $\Gamma$, and the norm $\|\Lambda\|_{L_{d/(s-1)}(\Omega)}$.
Proof. Since $\Lambda \in H^1(\Omega)$, by the embedding theorem we have
where the embedding constant $\tilde{c}(d,\Omega)$ depends on $d$ and $\Omega$ (and on $r$ in the case $d=2$).
If $d \leqslant 2s$, then it follows from (14.72) that the assumptions of Proposition 14.29 for $l=s-1$ and $\Upsilon=\Lambda$ are satisfied. Hence the operator $[\Lambda]$ is continuous from $H^{s-1}$ to $L_2$. The required estimate for its norm follows from (14.68), (14.73), and inequalities (6.14) and (6.15). This proves statement $1^\circ$.
Statement $2^\circ$ follows directly from Proposition 14.29. $\Box$
The following statement was obtained in [14], Proposition 9.3.
Proposition 14.31 ([14]). Let $\Lambda$ be a $\Gamma$-periodic solution of problem (6.8). Let $l=1$ for $d=1$, $l>1$ for $d=2$, and $l=d/2$ for $d \geqslant 3$. Then the operator $[\Lambda]$ is continuous from $H^l(\mathbb{R}^d;\mathbb{C}^m)$ to $H^1(\mathbb{R}^d;\mathbb{C}^n)$, and the norm $\|[\Lambda]\|_{H^l \to H^1}$ is controlled in terms of $d$, $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and the parameters of the lattice $\Gamma$, and for $d=2$ it also depends on $l$.
Proposition 14.31 implies the following result directly.
Corollary 14.32. Let $s \geqslant 2$. For $d\leqslant 2s-2$ the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$, and the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$ is controlled in terms of $d$, $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and the parameters of the lattice $\Gamma$.
Next, the following statement is true.
Proposition 14.33. Let $\Lambda$ be the $\Gamma$-periodic solution of problem (6.8). Let $l \geqslant 1$ and $d > 2l-2$. Suppose that $\Lambda \in L_{d/(l-1)}(\Omega)$. Then the operator $[\Lambda]$ is continuous from $H^l(\mathbb{R}^d;\mathbb{C}^m)$ to $H^1(\mathbb{R}^d;\mathbb{C}^n)$, and the norm $\|[\Lambda]\|_{H^l \to H^1}$ is controlled in terms of $d$, $l$, $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, the parameters of the lattice $\Gamma$, and the norm $\|\Lambda\|_{L_{d/(l-1)}(\Omega)}$.
For $l=2$ this statement was checked in [27], Lemma 8.7; the case of arbitrary $l\geqslant 1$ is considered similarly.
We present other sufficient conditions for the continuity of $[\Lambda]$ from $L_2$ to $L_2$ and from $H^1$ to $H^1$. (Note that the continuity of the operator $[\Lambda]$ from $L_2$ to $L_2$ implies its continuity from $H^{s-1}$ to $L_2$ for any $s \geqslant 1$, and the continuity of the operator $[\Lambda]$ from $H^1$ to $H^1$ implies its continuity from $H^{s-1}$ to $H^1$ for any $s \geqslant 2$.)
Proposition 14.34. Suppose that at least one of the following assumptions is satisfied:
(a) the dimension $d$ is arbitrary and $\widehat{\mathcal A}={\mathbf D}^* g({\mathbf x}){\mathbf D}$, where the matrix $g({\mathbf x})$ has real entries;
(b) the dimension $d$ is arbitrary and $g^0=\underline{g}$ (that is, relations (6.21) are fulfilled).
Then the operator $[\Lambda]$ is continuous from $L_2(\mathbb{R}^d;\mathbb{C}^m)$ to $L_2(\mathbb{R}^d;\mathbb{C}^n)$ and from $H^{1}(\mathbb{R}^d;\mathbb{C}^m)$ to $H^1(\mathbb{R}^d;\mathbb{C}^n)$. The norms $\|[\Lambda]\|_{L_2 \to L_2}$ and $\|[\Lambda]\|_{H^{1} \to H^1}$ are controlled in terms of $d$, $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and the parameters of the lattice $\Gamma$.
Proof. In case (a) it follows from Theorem 13.1 in [112], Chap. III, that $\Lambda \in L_\infty$ (together with an estimate for the norm $\|\Lambda\|_{L_\infty}$ in terms of $d$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, and $\Omega$). It remains to apply Corollaries 14.30, 14.32, and Proposition 14.33.
In the case where $g^0=\underline{g}$ the relation $\Lambda \in L_\infty$ (together with a suitable estimate for the norm $\|\Lambda\|_{L_\infty}$) was obtained in [9], Proposition 6.9. Again, we apply Corollaries 14.30 and 14.32 and Proposition 14.33. This completes the proof.
14.9. Special cases
We consider the special cases where $g^0=\overline{g}$ or $g^0=\underline{g}$.
Proposition 14.35. Suppose that $g^0=\overline{g}$, that is, relations (6.20) are satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constants $\widehat{\mathfrak{C}}_{8}(s)$, $\widehat{\mathfrak{C}}_{9}'(s)$, $\widehat{\mathfrak{C}}_{10}(s)$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $r_0$, and $s$.
Proof. If relations (6.20) are true, then $\Lambda({\mathbf x}) =0$. Then Condition 7.2 is satisfied, and the operator (14.11) takes the form $\widehat{G}_{0,\varepsilon}(\tau)=e^{-i\tau\widehat{\mathcal{A}}_\varepsilon} -e^{-i\tau\widehat{\mathcal{A}}^{\,0}}$. Therefore, estimate (14.74) follows directly from Corollaries 14.4 and 14.10; we have $\widehat{\mathfrak{C}}_{10}(s)=\widehat{\mathfrak{C}}_{2}(s)$ if $0\leqslant s \leqslant 2$ and $\widehat{\mathfrak{C}}_{10}(s)=\widehat{\mathfrak{C}}_{4}(s)$ if $2 < s \leqslant 4$. Estimate (14.75) follows from Corollary 14.16, and (14.76) follows from (14.75); we have $\widehat{\mathfrak{C}}_{9}'(s)= \|g\|_{L_\infty} \alpha_1^{1/2} \widehat{\mathfrak{C}}_{8}(s)$. This completes the proof.
Proposition 14.36. Suppose that $g^0=\underline{g}$, that is, relations (6.21) are satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ estimates (14.10), (14.55), and (14.65) are true, and estimate (14.66) takes the form
Proof. If relations (6.21) are valid, then $\widetilde{g}({\mathbf x})=g^0=\underline{g}$. According to statement $3^\circ$ of Proposition 6.4, Condition 7.2 is satisfied. By Corollary 14.4 estimate (14.10) is fulfilled. By statement $2^\circ$ of Proposition 14.34 results ‘without the smoothing operator’ can be used: Theorems 14.24 and 14.28 imply estimates (14.55), (14.65), and (14.66). This completes the proof.
15. Approximation for the sandwiched operator exponential $e^{-i\tau\mathcal{A}_\varepsilon}$
15.1. Approximation of the operator $f^\varepsilon e^{-i \tau \mathcal{A}_\varepsilon} (f^\varepsilon)^{-1}$ in the principal order
Now we proceed to the operator $\mathcal{A}_\varepsilon$ (see (14.2)). Let $\mathcal{A}^0$ be the operator (9.3). We put
where the operator ${\mathcal J}(\tau)$ is defined by (13.1).
Applying Theorems 13.1 and 13.2 and taking (15.2) into account, we obtain the following two theorems.
Theorem 15.1 ([25]). Let $\mathcal{A}_{\varepsilon}$ and $\mathcal{A}^0$ be the operators (14.2) and (9.3). Let ${\mathcal J}_{\varepsilon}(\tau)$ be the operator defined by (15.1). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_1$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 15.2 ([30]). Suppose that the assumptions of Theorem 15.1 are satisfied. Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 10.2 the constant ${\mathrm{C}}_2$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 10.4 this constant depends on the same parameters and on $n$ and $c^\circ$.
Previously, Theorem 15.1 was obtained in [25], Theorem 12.3, and Theorem 15.2 was proved in [30], Theorems 10.9 and 10.10.
By interpolation, Theorems 15.1 and 15.2 imply the following statements.
Corollary 15.3. Under the assumptions of Theorem 15.1, for $\tau \in \mathbb{R}$ and $\varepsilon>0$ we have
The operator (15.8) is bounded, and the operator (15.9) is in the general case defined on $H^3(\mathbb{R}^d;\mathbb{C}^n)$. Under Condition 10.4, the operator (15.9) is defined on $H^1(\mathbb{R}^d;\mathbb{C}^n)$.
Let ${\mathcal G}_{0}(\varepsilon^{-2} \tau)$ and ${\mathcal G}(\varepsilon^{-2} \tau)$ be the operators defined by (13.2) and (13.3). Applying the scaling transformation we obtain
Combining this with Theorems 13.3, 13.4, and 13.5 and taking into account that the operator $T_\varepsilon$ is unitary, we obtain the following statements directly.
Theorem 15.5. Let ${\mathcal G}_\varepsilon(\tau)$ be the operator defined by (15.9). For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_3$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 15.6. Let ${\mathcal G}_{0,\varepsilon}(\tau)$ be the operator defined by (15.8). Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_4$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Theorem 15.7. Let ${\mathcal G}_\varepsilon( \tau)$ be the operator defined by (15.9). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm{C}}_5$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^{\circ}$.
In a similar way Proposition 13.7 and relations (15.10) and (15.11) imply the following statement.
Proposition 15.8. Let ${\mathcal G}_{0,\varepsilon}(\tau)$ and ${\mathcal G}_\varepsilon(\tau)$ be the operators defined by (15.8) and (15.9). Then the following hold.
$1^\circ$. For $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm C}_6^\circ$ and ${\mathrm C}_6$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
$2^\circ$. Suppose that Condition 10.2 is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constant ${\mathrm C}_7$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
$3^\circ$. Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm C}_8^\circ$ and ${\mathrm C}_8$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and also on $n$ and ${c}^\circ$.
By interpolation, from Theorems 15.5–15.7 and Proposition 15.8 we deduce the following results.
Corollary 15.9. Under the assumptions of Theorem 15.5, for $\tau \in \mathbb{R}$ and $\varepsilon>0$ we have
By analogy with the proof of Theorem 14.12, using this identity and estimate (15.15) it is easy to deduce the following result from Theorem 13.8.
Theorem 15.12. Let ${\mathcal G}_{0,\varepsilon}(\tau)$ and $\Xi_\varepsilon(\tau)$ be the operators defined by (15.8) and (15.23). Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The constants ${\mathrm{C}}_{12}$, ${\mathrm{C}}_{13}$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$.
Next, by analogy with the proof of Theorem 14.13, applying Theorem 13.9 and using (15.24) and also (15.17) (under Condition 10.2) or (15.18) (under Condition 10.4), we obtain the following result.
Theorem 15.13. Let ${\mathcal G}_{0,\varepsilon}(\tau)$ and $\Xi_\varepsilon(\tau)$ be the operators defined by (15.8) and (15.23). Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under Condition 10.2, the constants ${\mathrm{C}}_{14}$, ${\mathrm{C}}_{15}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, and $r_0$. Under Condition 10.4, they depend on the same parameters, and also on $n$ and $c^\circ$.
By analogy with the proof of Proposition 14.14, it is easy to deduce the following statement from Proposition 13.10.
Proposition 15.14. Let ${\mathcal G}_{0,\varepsilon}(\tau)$ and $\Xi_\varepsilon(\tau)$ be the operators defined by (15.8) and (15.23). For $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constants ${\mathrm C}_{16}$ and ${\mathrm C}_{17}$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, and $r_1$.
By interpolation, from Theorems 15.12, 15.13 and Proposition 15.14 we obtain the following two statements.
Corollary 15.15. Under the assumptions of Theorem 15.12, for $\tau \in \mathbb{R}$ and $\varepsilon>0$ we have
Remark 15.17. (i) In the general case, that is, under the assumptions of Theorems 15.1, 15.5, and 15.12, we can consider large values of time $\tau=O(\varepsilon^{-\alpha})$, $0< \alpha < 1$, and obtain the qualified estimates
(ii) In the case of improvements of general results, that is, under the assumptions of Theorems 15.2, 15.6, 15.7, and 15.13, we can consider $\tau=O(\varepsilon^{-\alpha})$, $0< \alpha < 2$, and obtain the qualified estimates
15.4. The sharpness of the results of §§ 15.1–15.3
Applying theorems from § 13.4 we verify that the results of §§ 15.1–15.3 are sharp. We start with sharpness with respect to the type of the operator norm.
Using the scaling transformation, from Theorem 13.11 we deduce the following theorem, which demonstates the sharpness of general results (Theorems 15.1, 15.5, and 15.12).
Theorem 15.18. Suppose that Condition 11.1 is satisfied. Then the following hold.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that the estimate
holds for all sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [29], Theorem 13.12.
In a similar way, using the scaling transformation, Theorem 13.12 implies that the improved results (Theorems 15.2, 15.6, 15.7, and 15.13) are also sharp.
Theorem 15.19. Suppose that Condition 11.2 is satisfied.
$1^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 2$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (15.35) holds for sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 4$. Then there does not exist a constant $\mathcal{C}(\tau)$, such that estimate (15.36) holds for sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $0 \ne \tau \in \mathbb{R}$ and $0 \leqslant s < 3$. Then there does not exist a constant $\mathcal{C}(\tau)$ such that estimate (15.37) holds for sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.12.
We proceed to verifying the sharpness of the results with respect to the dependence of estimates on the parameter $\tau$.
Using the scaling transformation, Theorem 13.13 implies the following result, which confirms the sharpness of the general results (Theorems 15.1, 15.5, 15.12).
Theorem 15.20. Suppose that Condition 11.1 is satisfied.
$1^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau| = 0$ and estimate (15.35) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 6$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /\tau^2 = 0$ and estimate (15.36) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (15.37) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.13.
Finally, the sharpness of the improved results (Theorems 15.2, 15.6, 15.7, and 15.13) follows from Theorem 13.14 using the scaling transformation.
Theorem 15.21. Suppose that Condition 11.2 is satisfied. Then the following hold.
$1^\circ$. Let $s \geqslant 2$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) /|\tau|^{1/2} =0$ and estimate (15.35) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$2^\circ$. Let $s \geqslant 4$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau| = 0$ and estimate (15.36) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
$3^\circ$. Let $s \geqslant 3$. Then there does not exist a positive function $\mathcal{C}(\tau)$ such that $\lim_{\tau \to \infty} \mathcal{C}(\tau) / |\tau|^{1/2} =0$ and estimate (15.37) holds for $\tau \in \mathbb{R}$ and sufficiently small $\varepsilon > 0$.
Previously, statement $1^\circ$ was obtained in [30], Theorem 10.14.
15.5. On the possible removal of the smoothing operator $\Pi_\varepsilon$ from approximations
Now we consider the question about the possibile removal of the operator $\Pi_\varepsilon$ from approximations in the results of §§ 15.2 and 15.3. We put
Similarly to the proof of Theorem 14.23 it is easy to deduce the following statement from Corollary 15.9, relation (15.40), and Lemma 14.22. It should be taken into account that the matrix-valued functions $\Lambda_Q$ and $\Lambda$ differ by a constant term, and therefore they have the same multiplier properties (in the Sobolev spaces).
Theorem 15.22. Let ${\mathcal G}'_\varepsilon(\tau)$ be the operator defined by (15.39). Let $3\leqslant s\leqslant 6$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant ${\mathfrak C}'_3(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
In a similar way Corollaries 15.10 and 15.11 and Lemma 14.22 imply the following results.
Theorem 15.23. Let ${\mathcal G}'_{0,\varepsilon}(\tau)$ be the operator defined by (15.38). Suppose that Condition 10.2 is satisfied. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant ${\mathfrak C}'_4(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
Theorem 15.24. Let ${\mathcal G}'_\varepsilon(\tau)$ be the operator defined by (15.39). Suppose that Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constant ${\mathfrak C}'_5(s)$ depends on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, $n$, ${c}^{\circ}$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to L_2}$.
Now, we consider the question about the possible removal of the operator $\Pi_\varepsilon$ from approximations of the operator exponential in the ‘energy’ norm. We put
By analogy with the proof of Theorem 14.26, it is easy to deduce the following statement from Corollary 15.15 and Lemma 14.22.
Theorem 15.25. Let ${\mathcal G}'_{0,\varepsilon}(\tau)$ and $\Xi_\varepsilon'(\tau)$ be the operators defined by (15.38) and (15.42). Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
The constants ${\mathfrak C}'_6(s)$ and ${\mathfrak C}'_7(s)$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$.
In a similar way Corollary 15.16 and Lemma 14.22 imply the following result.
Theorem 15.26. Let ${\mathcal G}'_{0,\varepsilon}( \tau)$ and $\Xi_\varepsilon'(\tau)$ be the operators defined by (15.38) and (15.42). Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. Let $2 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. Then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
Under Condition 10.2, the constants ${\mathfrak C}'_{8}(s)$ and ${\mathfrak C}'_9(s)$ depend on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f\|_{L_\infty}$, $\|f^{-1}\|_{L_\infty}$, $r_0$, $s$, and also on the norm $\|[\Lambda]\|_{H^{s-1} \to H^1}$. Under Condition 10.4, these constants depend on the same parameters and on $n$, ${c}^{\circ}$.
Recall that some cases where one of the conditions on the operator $[\Lambda]$ (from Theorems 15.22–15.26) is fulfilled are listed in Corollaries 14.30 and 14.32 and Propositions 14.33 and 14.34.
15.6. Special cases
We consider two special cases where $g^0=\overline{g}$ or $g^0=\underline{g}$.
Proposition 15.27. Suppose that $g^0=\overline{g}$, that is, relations (6.20) are satisfied. Let ${\mathcal J}_\varepsilon(\tau)$ be the operator (15.1). Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
The constants ${\mathfrak{C}}_{8}(s)$, ${\mathfrak{C}}_{9}'(s)$, and ${\mathfrak{C}}_{10}(s)$ depend only on $\alpha_0$, $\alpha_1$, $\|g\|_{L_\infty}$, $\|g^{-1}\|_{L_\infty}$, $\|f \|_{L_\infty}$, $\| f^{-1}\|_{L_\infty}$, $r_0$, and $s$.
Proof. It follows from (6.20) that $\Lambda({\mathbf x}) =0$ and $\Lambda_Q({\mathbf x})=0$. Then Condition 10.2 is satisfied, and the operator (15.8) takes the form ${\mathcal G}_{0,\varepsilon}(\tau)={\mathcal J}_\varepsilon(\tau)$. Therefore, estimate (15.45) follows directly from Corollaries 15.4 and 15.10; moreover, ${\mathfrak{C}}_{10}(s)={\mathfrak{C}}_{2}(s)$ if $0\leqslant s \leqslant 2$ and ${\mathfrak{C}}_{10}(s)={\mathfrak{C}}_{4}(s)$ if $2 < s \leqslant 4$. Estimate (15.46) follows from Corollary 15.16, and (15.47) follows from (15.46); moreover, ${\mathfrak{C}}_{9}'(s)=\|g\|_{L_\infty}\alpha_1^{1/2}{\mathfrak{C}}_{8}(s)$. This completes the proof.
Proposition 15.28. Suppose that $g^0=\underline{g}$, that is, relations (6.21) are satisfied. Then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ estimates (15.5), (15.41), and (15.43) are true, and estimate (15.44) takes the form
Proof. If relations (6.21) are satisfied, then $\widetilde{g}({\mathbf x})=g^0=\underline{g}$. By Corollary 15.3, estimate (15.5) is fulfilled. By statement $2^\circ$ of Proposition 14.34, results ‘without smoothing’ can be applied: Theorems 15.22 and 15.25 imply estimates (15.41), (15.43), and (15.44). This completes the proof.
16. Homogenization of the Cauchy problem for the Schrödinger-type equation
16.1. The Cauchy problem with the operator $\widehat{\mathcal{A}}_\varepsilon$. The principal term of the approximation of the solution
Let $\check{\mathbf{u}}_\varepsilon(\mathbf{x},\tau)$ be the solution of the following Cauchy problem:
Theorem 16.1. Let $\check{\mathbf{u}}_\varepsilon$ be the solution of problem (16.1), and let ${\mathbf{u}}_0$ be the solution of the homogenized problem (16.3).
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
Proof. Estimate (16.5) follows directly from Corollary 14.3 and representations (16.2) and (16.4). By the Banach–Steinhaus theorem statement $1^\circ$ implies statement $2^\circ$. This completes the proof.
Statement $1^\circ$ of Theorem 16.1 can be improved under certain additional assumptions. Corollary 14.4 implies the following result.
Theorem 16.2. Suppose that $\check{\mathbf{u}}_\varepsilon$ is the solution of problem (16.1) and $\mathbf{u}_0$ is the solution of the homogenized problem (16.3). Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $0 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
16.2. The Cauchy problem with the operator $\widehat{\mathcal{A}}_\varepsilon$ and initial data from a special class. More accurate approximation of the solution
The results with correctors (see §§ 14.4 and 14.5) can be applied to the Cauchy problem with data $\boldsymbol{\phi}_\varepsilon$ from a special class.
Let $\mathbf{u}_\varepsilon(\mathbf{x},\tau)$ be the solution of the following Cauchy problem:
This easily follows from the relation $\Pi_\varepsilon[\Lambda^\varepsilon]b({\mathbf D})\Pi_\varepsilon=0$, which, by the scaling transformation and the direct integral decomposition, is reduced to the equality $\widehat{P}[\Lambda]b(\mathbf k)\widehat{P}=0$. The latter follows from the condition $ \int_\Omega \Lambda({\mathbf x})\,d{\mathbf x}=0$. Hence we have $\boldsymbol{\phi}=(I-\varepsilon\Lambda^\varepsilon b({\mathbf D}) \Pi_\varepsilon)\boldsymbol{\phi}_\varepsilon$.
Let $\mathbf{u}_0(\mathbf{x},\tau)$ be the solution of the previous homogenized problem (16.3), and let $\mathbf{w}_0(\mathbf{x},\tau)$ be the solution of the problem
Theorem 16.4. Let ${\mathbf{u}}_\varepsilon$ be the solution of problem (16.6) with initial data of the form (16.7), and let ${\mathbf p}_\varepsilon$ be defined by (16.11). Let $\mathbf{u}_0$ be the solution of the homogenized problem (16.3), and let ${\mathbf v}_\varepsilon$ be defined by (16.12). Let $\mathbf{w}_0$ be the solution of problem (16.9).
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$2^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$3^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $1 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
The second term is estimated by using the scaling transformation and the direct integral decomposition. By (7.9) and (7.12), for $0\leqslant s \leqslant 3$ we have
where $\widehat{G}_{0,\varepsilon}(\tau)$ and $\widehat{\Xi}_{\varepsilon}(\tau)$ are the operators defined by (14.11) and (14.26). In combination with Corollary 14.15, this implies estimates (16.15) and (16.16). $\Box$
Under the additional assumptions, the results of Theorem 16.4 can be improved. By analogy with the proof of Theorem 16.4, from Corollaries 14.4, 14.10, 14.11, and 14.16 we deduce the following result.
Theorem 16.5. Suppose that the assumptions of Theorem 16.4 are satisfied. Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied.
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $0 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$2^\circ$. Under Condition 7.2, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$3^\circ$. Under Condition 7.4, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$4^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $1 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
16.3. The case where the smoothing operator can be removed
Now we consider the case where the conditions on $\Lambda$ are satisfied, allowing us to remove the smoothing operator $\Pi_\varepsilon$ from approximations; see § 14.8. Assuming that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$, where $s \geqslant 1$, we consider the following Cauchy problem:
Let $\mathbf{u}_0(\mathbf{x}, \tau)$ be the solution of the previous homogenized problem (16.3), and let $\mathbf{w}_0(\mathbf{x},\tau)$ be the solution of problem (16.9). We put
Corollary 14.3 and Theorems 14.23 and 14.26 imply the following result.
Theorem 16.6. Let $\tilde{\mathbf{u}}_\varepsilon$ be the solution of problem (16.20) with initial data of the form (16.21), and let $\tilde{{\mathbf p}}_\varepsilon$ be defined by (16.23). Suppose that $\mathbf{u}_0$ is the solution of the homogenized problem (16.3) and $\tilde{\mathbf v}_\varepsilon$ is defined by (16.24). Let $\mathbf{w}_0$ be the solution of problem (16.9). Then the following hold.
$1^\circ$. Let $1 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$2^\circ$. Let $3 \leqslant s \leqslant 6$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$3^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0<\varepsilon \leqslant 1$ we have
where $\widehat{G}'_{0,\varepsilon}(\tau)$ and $\widehat{\Xi}'_{\varepsilon}(\tau)$ are the operators defined by (14.50) and (14.56), respectively. Combining this with Theorem 14.26, we obtain estimates (16.27) and (16.28). $\Box$
Under certain additional assumptions the results of Theorem 16.6 can be improved. By analogy with the proof of Theorem 16.6, from Corollary 14.4 and Theorems 14.24, 14.25, and 14.28 we deduce the following result.
Theorem 16.7. Suppose that the assumptions of Theorem 16.6 are satisfied. Suppose that Condition 7.2 or Condition 7.4 (or the more restrictive Condition 7.5) is satisfied. Then the following hold.
$1^\circ$. Let $1 \leqslant s \leqslant 2$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$2^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Under Condition 7.2, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$3^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Under Condition 7.4, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$4^\circ$. Let $2 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
Theorem 16.8. Let $\check{\mathbf u}_\varepsilon$ be the solution of problem (16.31), and let ${\mathbf u}_0$ be the solution of the homogenized problem (16.33).
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon >0$ we have
Proof. Estimate (16.35) follows directly from Corollary 15.3 and representations (16.32) and (16.34). Statement $2^\circ$ follows from statement $1^\circ$ by the Banach– Steinhaus theorem. $\Box$
Statement $1^\circ$ of Theorem 16.8 can be improved under additional assumptions. From Corollary 15.4 we deduce the following result.
Theorem 16.9. Let $\check{\mathbf u}_\varepsilon$ be the solution of problem (16.31), and let ${\mathbf u}_0$ be the solution of the homogenized problem (16.33). Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $0 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
16.5. The Cauchy problem with the operator ${\mathcal{A}}_\varepsilon$ and initial data from a special class. More accurate approximation of the solution
Results with correctors (see §§ 15.2, 15.3) can be applied to the Cauchy problem with data $\boldsymbol{\phi}_\varepsilon$ from a special class.
Let ${\mathbf u}_\varepsilon(\mathbf{x},\tau)$ be the solution of the following Cauchy problem:
Let ${\mathbf u}_0(\mathbf{x},\tau)$ be the solution of the previous homogenized problem (16.33). Let ${\mathbf w}_0(\mathbf{x},\tau)$ be the solution of the problem
Theorem 16.10. Let ${\mathbf u}_\varepsilon$ be the solution of problem (16.36) with the initial data of the form (16.37), and let ${\mathbf p}_\varepsilon$ be defined by (16.41). Let ${\mathbf u}_0$ be the solution of the homogenized problem (16.33), and let ${\mathbf v}_\varepsilon$ be defined by (16.42). Let ${\mathbf w}_0$ be the solution of problem (16.39).
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$2^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$3^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $1 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Under some additional assumptions, the results of Theorem 16.10 can be improved. Corollaries 15.4, 15.10, 15.11, and 15.16 imply the following result.
Theorem 16.11. Suppose that the assumptions of Theorem 16.10 are satisfied. Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied.
$1^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $0 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$2^\circ$. Under Condition 10.2, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$3^\circ$. Under Condition 10.4, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$4^\circ$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, where $1 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
16.6. The case where the smoothing operator can be removed
Now we consider the case where the conditions on $\Lambda$ are satisfied, allowing us to remove the smoothing operator $\Pi_\varepsilon$ from approximations; see § 15.5. Assuming that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$, where $s \geqslant 1$, we consider the following Cauchy problem:
Let $\mathbf{u}_0 (\mathbf{x},\tau)$ be the solution of the previous homogenized problem (16.33), and let $\mathbf{w}_0(\mathbf{x},\tau)$ be the solution of problem (16.39). We put
By analogy with the proof of Theorem 16.6, from Corollary 15.3 and Theorems 15.22 and 15.25 we deduce the following result.
Theorem 16.12. Let $\tilde{\mathbf{u}}_\varepsilon$ be the solution of problem (16.43) with initial data of the form (16.44), and let $\tilde{{\mathbf p}}_\varepsilon$ be defined by (16.46). Let $\mathbf{u}_0$ be the solution of the homogenized problem (16.33), and let $\tilde{\mathbf v}_\varepsilon$ be defined by (16.47). Let $\mathbf{w}_0$ be the solution of problem (16.39).
$1^\circ$. Let $1 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$2^\circ$. Let $3 \leqslant s \leqslant 6$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$3^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
Under some additional assumptions, the results of Theorem 16.12 can be improved. Corollary 15.4 and Theorems 15.23, 15.24, and 15.26 imply the following result.
Theorem 16.13. Suppose that the assumptions of Theorem 16.12 are satisfied. Suppose that Condition 10.2 or Condition 10.4 (or the more restrictive Condition 10.5) is satisfied.
$1^\circ$. Let $1 \leqslant s \leqslant 2$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d;\mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$2^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Under Condition 10.2, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$3^\circ$. Let $2 \leqslant s \leqslant 4$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$. Under Condition 10.4, if $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$4^\circ$. Let $2 \leqslant s \leqslant 3$. Suppose that the operator $[\Lambda]$ is continuous from $H^{s-1}(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$. If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d; \mathbb{C}^n)$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
Here $g({\mathbf x})$ is a $\Gamma$-periodic Hermitian $d \times d $ matrix-valued function such that $g({\mathbf x}) \!>\!0$ and $g,g^{-1} \in L_\infty$. The operator (17.1) is a particular case of the operator (6.1). In this case we have $n=1$, $m=d$, and $b({\mathbf D})={\mathbf D}$. Obviously, condition (5.7) is satisfied for $\alpha_0=\alpha_1=1$. According to (6.17), the effective operator for the operator (17.1) is given by $\widehat{\mathcal A}^0= {\mathbf D}^* g^0{\mathbf D}=-\operatorname{div} g^0 \nabla$. By (6.11) and (6.12) the effective matrix $g^0$ is defined as follows. Let ${\mathbf e}_1,\dots,{\mathbf e}_d$ be the standard orthonormal basis in $\mathbb{R}^d$. Let $\Phi_j \in \widetilde{H}^1(\Omega)$ be the weak $\Gamma$-periodic solution of the problem
Then $\Lambda({\mathbf x})$ is a row: $\Lambda({\mathbf x})= i\bigl(\Phi_1({\mathbf x}),\dots,\Phi_d({\mathbf x})\bigr)$, and $\widetilde{g}({\mathbf x})$ is the $d\times d$ matrix with columns $\widetilde{\mathbf g}_j({\mathbf x})=g({\mathbf x}) \bigl(\nabla \Phi_j({\mathbf x})+{\mathbf e}_j\bigr)$, $j=1,\dots,d$. The effective matrix is defined by $g^0=|\Omega|^{-1} \int_\Omega \widetilde{g}({\mathbf x})\,d{\mathbf x}$. In the case where $d=1$ we have $m=n=1$, so that $g^0=\underline{g}$.
The first eigenvalue of the operator $\widehat{\mathcal A}(\mathbf k)=\widehat{A}(t,\boldsymbol{\theta})$ admits a power series expansion:
where $\widehat{\gamma}(\boldsymbol{\theta})= \langle g^0\boldsymbol{\theta},\boldsymbol{\theta}\rangle$. Since $n=1$, $\widehat{N}(\boldsymbol{\theta})=\widehat{N}_0(\boldsymbol{\theta})$ is the operator of multiplication by $\widehat{\mu}(\boldsymbol{\theta})$.
If $g({\mathbf x})$ is a symmetric matrix with real entries, then from statement $1^\circ$ of Proposition 6.4 it follows that $\widehat{N}(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$, that is, Condition 7.2 is satisfied. By Proposition 14.34 the operator $[\Lambda]$ is continuous from $L_2(\mathbb{R}^d)$ to $L_2(\mathbb{R}^d)$ and from $H^1(\mathbb{R}^d)$ to $H^1(\mathbb{R}^d)$.
If $g({\mathbf x})$ is a Hermitian matrix with complex entries, then in general the operator $\widehat{N}(\boldsymbol{\theta})$ is non-trivial; see an example in [9], § 10.4. A calculation (see [9], § 10.3) shows that
Now we describe the operator $\widehat{\mathcal N}^{(1)}(\boldsymbol{\theta})$ which acts as multiplication by $\widehat{\nu}(\boldsymbol{\theta})$. Suppose that $\Psi_{jl}({\mathbf x})$ is the $\Gamma$-periodic solution of the problem
Remark 17.1. As shown in [30], Lemma 12.2, if $d=1$ and $g(x) \ne \operatorname{const}$, then $\widehat{\nu}(1)=\widehat{\nu}(-1) \ne 0$. Therefore, the author believes that, as a rule, in the multidimensional case $\widehat{\nu}(\boldsymbol{\theta}) \ne 0$.
17.2. Homogenization
In the general case we apply Theorems 14.1, 14.5, and 14.12 and Corollaries 14.3, 14.9, and 14.15 to the operator
In the ‘real’ case, we apply the ‘improved’ results (Theorems 14.2, 14.6, and 14.13 and Corollaries 14.4, 14.10, and 14.16) and also the results ‘without smoothing’ (Theorems 14.24 and 14.28).
Applying Theorem 16.4 in the general case and Theorem 16.5 in the ‘real’ case, we obtain the following result.
Proposition 17.2. Let ${u}_\varepsilon$ be the solution of problem (17.4), and let ${u}_0$ be the solution of the homogenized problem (17.5). Let $w_0$ be the solution of problem (17.6). Let $v_\varepsilon$ be defined by (17.7).
$1^\circ$. If $\phi \in H^{s}(\mathbb{R}^d)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Applying Theorem 16.7 and using the fact that in the ‘real’ case Condition 7.2 is satisfied, we obtain the following result.
Proposition 17.3. Suppose that $g({\mathbf x})$ is a symmetric matrix with real entries. Let $\tilde{u}_\varepsilon$ be the solution of problem (17.8), and let ${u}_0$ be the solution of the homogenized problem (17.5). Let $\tilde{v}_\varepsilon$ be defined by (17.9).
If $\phi \in H^{s}(\mathbb{R}^d)$, where $1 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
$$
\begin{equation*}
\|\tilde{u}_\varepsilon(\,{\cdot}\,,\tau)- {u}_0(\,{\cdot}\,,\tau)\|_{L_2(\mathbb{R}^d)} \leqslant C (s) (1+|\tau|)^{s/4} \varepsilon^{s/2} \|{\phi}\|_{H^{s}(\mathbb{R}^d)}.
\end{equation*}
\notag
$$
If ${\phi} \in H^{s}(\mathbb{R}^d)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^d)$, where $2 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
where the metric $\check{g}({\mathbf x})$ and the potential $V({\mathbf x})$ are $\Gamma$-periodic. It is assumed that $\check{g}({\mathbf x})$ is a symmetric $ d \times d $ matrix-valued function with real entries such that $\check{g},\check{g}^{-1} \in L_\infty$ and $\check{g}({\mathbf x}) >0$. The potential $V({\mathbf x})$ is assumed to be real-valued and such that
$$
\begin{equation}
V \in L_q(\Omega), \qquad q=1 \ \ \text{for}\ d=1, \quad 2q > d \ \ \text{for}\ d \geqslant 2.
\end{equation}
\tag{18.2}
$$
More precisely, $\mathcal H$ is a selfadjoint operator in $L_2(\mathbb{R}^d)$ corresponding to the quadratic form
Under the above assumptions, the form (18.3) is lower semibounded and closed. Adding an appropriate constant to $V({\mathbf x})$, we assume that the point $\lambda_0=0$ is the lower edge of the spectrum of the operator $\mathcal H$.
Then the equation ${\mathbf D}^* \check{g}({\mathbf x}){\mathbf D} \omega({\mathbf x})+ V({\mathbf x}) \omega({\mathbf x})=0$ has a (weak) positive $\Gamma$-periodic solution $\omega \in \widetilde{H}^1(\Omega)$. We have $\omega \in C^\sigma$ for some $\sigma >0$. Moreover, the function $\omega$ is a multiplier in the classes $H^1(\mathbb{R}^d)$ and $\widetilde{H}^1(\Omega)$. We fix the solution $\omega$ by the normalization condition
Thus, the operator $\mathcal H$ takes the form (5.10), where $n=1$, $m=d$, $b({\mathbf D})={\mathbf D}$, $g=\omega^2\check{g}$, and $f=\omega^{-1}$.
Remark 18.1. Expression (18.5) can be viewed as the definition of the operator $\mathcal H$, assuming that $\omega$ is an arbitrary $\Gamma$-periodic function satisfying conditions $\omega,\omega^{-1}\in L_\infty$, $\omega({\mathbf x}) >0$, and also condition (18.4). We can return to the form (18.1) by putting $V=-\omega^{-1}({\mathbf D}^*\check{g}{\mathbf D}\omega)$. The corresponding potential $V({\mathbf x})$ can be a singular distribution.
The operator (18.5) is related to the operator (17.1) (for $g=\omega^2 \check{g}$) by the identity ${\mathcal H}=\omega^{-1} \widehat{\mathcal A} \omega^{-1}$. Let $g^0$ be the effective matrix for the operator (17.1), which we found in § 17.1. Now the function $Q=(f f^*)^{-1}$ takes the form $Q({\mathbf x})=\omega^2({\mathbf x})$. By condition (18.4) we have $\overline{Q}=1$, hence $f_0=(\overline{Q})^{-1/2}=1$. The operator (9.3) takes the form
Thus, ${\mathcal H}^0$ coincides with the effective operator for the operator $\widehat{\mathcal A}={\mathbf D}^*g {\mathbf D}= {\mathbf D}^* \omega^2 \check{g} {\mathbf D}$.
The matrix $\Lambda_Q({\mathbf x})$ is the row $\Lambda_Q({\mathbf x}) =i\bigl(\Phi_{1,Q}({\mathbf x}),\dots,\Phi_{d,Q}({\mathbf x})\bigr)$, where $\Phi_{j,Q} \in \widetilde{H}^1(\Omega)$ is the weak $\Gamma$-periodic solution of the problem
By statement $1^\circ$ of Proposition 9.1 we have $\widehat{N}_Q(\boldsymbol{\theta})=0$ for any $\boldsymbol{\theta} \in \mathbb{S}^{d-1}$, that is, Condition 10.2 is satisfied. The first eigenvalue of the operator ${\mathcal H}(\mathbf k)$ admits the power series expansion
where $\gamma(\boldsymbol{\theta})= \langle g^0 \boldsymbol{\theta}, \boldsymbol{\theta}\rangle$. (Note that in quantum mechanics the matrix $(2g^0)^{-1}$ is called the tensor of the effective masses.)
18.2. Homogenization of the non-stationary Schrödinger equation
Since there is a large factor $\varepsilon^{-2}$ in front of the rapidy oscillating function $V^\varepsilon$, the second term in (18.7) is called a ‘strongly singular potential’. We can apply the ‘improved’ results (Theorems 15.2, 15.6, and 15.13 and Corollaries 15.4, 15.10, and 15.16) and also the results ‘without smoothing’ (Theorems 15.23 and 15.26) to the operator ${\mathcal H}_\varepsilon$.
Applying Theorem 16.11 we obtain the following result.
Proposition 18.2. Let ${u}_\varepsilon$ be the solution of problem (18.8), and let ${u}_0$ be the solution of the homogenized problem (18.9). Let $v_\varepsilon$ be defined by (18.10).
If $\phi \in H^{s}(\mathbb{R}^d)$, where $0 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$$
\begin{equation*}
\|(\omega^\varepsilon)^{-1}{u}_\varepsilon(\,{\cdot}\,,\tau)- {u}_0 (\,{\cdot}\,, \tau)\|_{L_2 (\mathbb{R}^d)} \leqslant C (s) (1+|\tau|)^{s/4} \varepsilon^{s/2} \| {\phi} \|_{H^{s}(\mathbb{R}^d)}.
\end{equation*}
\notag
$$
If ${\phi} \in H^{s}(\mathbb{R}^d)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Applying Theorem 16.13 we obtain the following result.
Proposition 18.3. Let $\tilde{u}_\varepsilon$ be the solution of problem (18.11) and let ${u}_0$ be the solution of the homogenized problem (18.9). Let $\tilde{v}_\varepsilon$ be defined by (18.12).
If $\phi \in H^{s}(\mathbb{R}^d)$, where $1 \leqslant s \leqslant 2$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
$$
\begin{equation*}
\|(\omega^\varepsilon)^{-1} \tilde{u}_\varepsilon(\,{\cdot}\,, \tau)- {u}_0 (\,{\cdot}\,, \tau)\|_{L_2 (\mathbb{R}^d)} \leqslant C (s) (1+|\tau|)^{s/4} \varepsilon^{s/2}\| {\phi} \|_{H^{s}(\mathbb{R}^d)}.
\end{equation*}
\notag
$$
If ${\phi} \in H^{s}(\mathbb{R}^d)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $0< \varepsilon \leqslant 1$ we have
where the metric $\check{g}({\mathbf x})$, the magnetic potential $\mathbf{A}({\mathbf x})$, and the electric potential $V({\mathbf x})$ are $\Gamma$-periodic. It is assumed that $\check{g}({\mathbf x})$ is a symmetric $d \times d $ matrix-valued function with real entries such that $\check{g},\check{g}^{-1} \!\in L_\infty$ and $\check{g}({\mathbf x})> 0$. If $d \geqslant 3$, then we suppose in addition that $\check{g} \in C^\sigma$ for some $\sigma$, $0< \sigma <1$. The $\mathbb{R}^d$-valued potential $\mathbf{A}({\mathbf x})$ and the real-valued potential $V({\mathbf x})$ are subject to the conditions
$$
\begin{equation*}
\mathbf{A} \in L_{2q}(\Omega),\quad V \in L_q(\Omega)\quad\text{and} \quad 2q > d.
\end{equation*}
\notag
$$
The precise definition of a selfadjoint operator $\mathcal M$ in $L_2(\mathbb{R}^d)$ is given in terms of the corresponding quadratic form. Adding an appropriate constant to $V({\mathbf x})$, we assume that the point $\lambda_0=0$ is the lower edge of the spectrum of the operator $\mathcal M$.
According to [113], under the above assumptions and for a sufficiently small potential $\mathbf A$ (in the $L_{2q}(\Omega)$-norm), the operator $\mathcal M$ admits a suitable factorization. To describe it we consider the family of operators ${\mathcal M}(\mathbf k)$ in $L_2(\Omega)$ arising in the direct integral decomposition of the operator $\mathcal M$. The condition
means that for some $\mathbf k_0 \in \widetilde{\Omega}$ the point $\lambda_0 =0$ is an eigenvalue of the operator ${\mathcal M}(\mathbf k_0)$. If the potential $\mathbf{A}$ is sufficiently small, then the point $\mathbf k_0$ is unique and $\lambda_0=0$ is a simple eigenvalue of ${\mathcal M}(\mathbf k_0)$. Let $\eta({\mathbf x})$ be the corresponding eigenfunction satisfying the normalization condition $ \int_\Omega|\eta({\mathbf x})|^2\,d{\mathbf x}=|\Omega|$ (the choice of a phase factor does not matter). Then $\eta \in \widetilde{H}^1(\Omega)$ and $\eta, \eta^{-1} \in L_\infty$. As shown in [113], the function $\eta({\mathbf x})$ is a multiplier in the classes $H^1(\mathbb{R}^d)$ and $\widetilde{H}^1(\Omega)$. We put
Clearly, the coefficients of the operator $\widetilde{\mathcal M}$ are periodic. By [113], Theorems 2.7 and 2.8, if the norm $\|{\mathbf A}\|_{L_{2q}(\Omega)}$ is sufficiently small, then the operator $\widetilde{\mathcal M}$ admits the following factorization:
As shown in [113], the matrix (18.14) is positive definite and such that $g,g^{-1} \in L_\infty$.
Thus, the operator (18.13) takes the form (5.10), where $n=1$, $m=d$, $b({\mathbf D})={\mathbf D}$, $g$ is defined by (18.14) and (18.15), and $f=\eta^{-1}$. Let $g^0$ be the effective matrix of the operator $\widehat{\mathcal A}={\mathbf D}^*g {\mathbf D}$. Next, the function $Q=(f f^*)^{-1}$ takes the form $Q({\mathbf x})=|\eta({\mathbf x})|^2$. By the normalization condition we have $\overline{Q}=1$, hence $f_0=1$. Now the role of the operator (9.3) is played by $\widetilde{\mathcal M}^0={\mathbf D}^* g^0 {\mathbf D}$. Let $\lambda(t,\boldsymbol{\theta})$ be the first eigenvalue of the operator $\widetilde{\mathcal M}(\mathbf k)$. The following power series expansion is fulfilled:
Let us describe the operator $\widehat{N}_Q(\boldsymbol{\theta})$. Since $n=1$ and $\overline{Q}=1$, the operator $\widehat{N}_Q(\boldsymbol{\theta})= \widehat{N}_{0,Q}(\boldsymbol{\theta})$ acts as multiplication by $\mu(\boldsymbol{\theta})$. A calculation shows that
where the coefficients $a_{jlk}$ are defined by (17.3) (see [29], § 15.4). In the general case ${\mu}(\boldsymbol{\theta})$ is not equal to zero. The third-order operator $\widehat{N}_Q({\mathbf D})$ takes the form
Note that expression (18.17) contains the large factors $\varepsilon^{-1}$ and $\varepsilon^{-2}$ in front of the potentials ${\mathbf A}^\varepsilon$ and $V^\varepsilon$, respectively. We can apply general results (Theorems 15.1, 15.5, and 15.12 and Corollaries 15.3, 15.9, and 15.15) to the operator $\widetilde{\mathcal M}_\varepsilon$.
Let $\check{u}_\varepsilon$ be the solution of the Cauchy problem for the non-stationary magnetic Schrödinger equation:
Then the function ${u}_\varepsilon({\mathbf x},\tau)=e^{-i\varepsilon^{-1} \langle \mathbf k_0,{\mathbf x}\rangle} \check{u}_\varepsilon({\mathbf x},\tau)$ is the solution of a problem of the form (16.36):
Here $\widehat{N}_Q({\mathbf D})$ is the operator defined by (18.16).
Applying Theorem 16.10 we obtain the following result.
Proposition 18.4. Suppose that $\check{u}_\varepsilon$ is the solution of problem (18.18) and ${u}_0$ is the solution of the homogenized problem (18.19). Let $v_\varepsilon$ be defined by (18.20), and let $w_0$ be the solution of problem (18.21).
If $\phi \in H^{s}(\mathbb{R}^d)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$$
\begin{equation*}
\|(\eta^\varepsilon)^{-1}e^{-i\varepsilon^{-1} \langle \mathbf k_0,\,{\cdot}\,\rangle} \check{u}_\varepsilon (\,{\cdot}\,,\tau)-{u}_0 (\,{\cdot}\,,\tau)\|_{L_2 (\mathbb{R}^d)} \leqslant C (s) (1+|\tau|)^{s/3} \varepsilon^{s/3}\|{\phi}\|_{H^{s}(\mathbb{R}^d)}.
\end{equation*}
\notag
$$
If ${\phi} \in H^{s}(\mathbb{R}^d)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
19. Applications of the general results: the two-dimensional Pauli equation
19.1. Definition of the two-dimensional Pauli operator. Factorization
(See [7], Chap. 6, § 2.1.) Let $d=2$. Suppose that the magnetic potential is given by the vector-valued function $\mathbf{A}({\mathbf x})=\{ A_1({\mathbf x}),A_2({\mathbf x})\}$, where the $A_j({\mathbf x})$, $j=1,2$, are real-valued $\Gamma$-periodic functions in $\mathbb{R}^2$ such that
which is closed in $L_2(\mathbb{R}^2;\mathbb{C}^2)$. If the vector-valued function $\mathbf{A}({\mathbf x})$ is Lipschitz, then the blocks $P_\pm$ of the operator (19.2) can be written as
As checked in [7], Chap. 6, § 2.1, we have $\varphi \in C^\sigma$ for $\sigma=1-2 \rho^{-1}$. We put $\omega_\pm({\mathbf x}):=e^{\pm \varphi({\mathbf x})}$. The operator $\mathcal P$ admits the following factorization:
Remark 19.1. (i) We can consider expressions (19.4) and (19.5) as the definitions of the operators $\mathcal P$ and $P_\pm$, respectively, by assuming that $\omega_\pm({\mathbf x})$ are arbitrary $\Gamma$-periodic functions satisfying the conditions
(ii) The operators $P_+$ and $P_-$ are unitarily equivalent. Moreover, the operators $P_+(\mathbf k)$ and $P_-(\mathbf k)$ acting in $L_2(\Omega)$ are also unitarily equivalent.
19.2. Effective characteristics of the operators $P_\pm$. Homogenization
The operator $P_\pm$ is of the form (5.10), where $m=n=1$, $b({\mathbf D})=D_1 \mp i D_2$, $g({\mathbf x})=\omega_\pm^2({\mathbf x})$, and $f({\mathbf x})=\omega_\mp({\mathbf x})$. The role of the operator $\widehat{\mathcal A}$ for $P_\pm$ is played by $\widehat{\mathcal A}_\pm=(D_1 \pm i D_2)\omega_\pm^2(D_1 \mp i D_2)$.
The role of the function $\Lambda({\mathbf x})$ for the operator $P_\pm$ is played by $\Lambda_\pm({\mathbf x})$, which is a $\Gamma$-periodic solution of the problem
$$
\begin{equation*}
(D_1 \pm i D_2)\omega_\pm^2({\mathbf x}) \bigl((D_1 \mp i D_2)\Lambda_\pm({\mathbf x})+1\bigr)=0,\qquad \int_\Omega \Lambda_\pm({\mathbf x}) \,d{\mathbf x}=0.
\end{equation*}
\notag
$$
Then the function $\widetilde{g}_\pm({\mathbf x}):=\omega_\pm^2({\mathbf x}) \bigl((D_1 \mp i D_2) \Lambda_\pm({\mathbf x}) +1\bigr)$ is constant. The effective constant $g^0_\pm$ is equal to the mean value of the function $\widetilde{g}_\pm({\mathbf x})$. Consequently,
(This is consistent with Proposition 6.1: in the case where $m=n$ we have $g^0=\underline{g}$.) Thus, $\Lambda_\pm({\mathbf x})$ is the $\Gamma$-periodic solution of the problem
The role of $Q({\mathbf x})$ for the operator $P_\pm$ is played by the function $Q_\pm({\mathbf x})=\omega_\pm^2({\mathbf x})$. Then $\overline{Q_\pm}=(g^0_\mp)^{-1}$. The role of $f_0$ is played by the constant $(\overline{Q_\pm})^{-1/2}= (g^0_\mp)^{1/2}=\omega_{\mp,0}$. Next, the role of the operator ${\mathcal A}^0$ for $P_\pm$ is played by the operator $P^0_\pm$, where
Let $\lambda_\pm(t, \boldsymbol{\theta})$ be the first eigenvalue of the operator $P_\pm(\mathbf k)$. Consider the corresponding power series expansion:
Since the operators $P_+(\mathbf k)$ and $P_-(\mathbf k)$ are unitarily equivalent, it follows that $\lambda_+(t,\boldsymbol{\theta})=\lambda_-(t,\boldsymbol{\theta})$, and also $\gamma_+(\boldsymbol{\theta})=\gamma_-(\boldsymbol{\theta})$ and $\mu_+(\boldsymbol{\theta})=\mu_-(\boldsymbol{\theta})$. As shown in [7], Chap. 6, § 2, the numbers $\gamma_\pm(\boldsymbol{\theta})$ do not depend on $\boldsymbol{\theta}$ and are given by $\gamma_+(\boldsymbol{\theta})=\gamma_-(\boldsymbol{\theta})=\gamma$, where $\gamma$ is defined by (19.8).
Similarly to (19.7), the role of $\Lambda_Q$ for the operator $P_\pm$ is played by $\Lambda_{Q,\pm}({\mathbf x})$, which is a $\Gamma$-periodic solution of the problem
Now, we describe the operator $\widehat{N}_{Q,\pm}(\boldsymbol{\theta})$ which plays the role of $\widehat{N}_{Q}(\boldsymbol{\theta})$ for $P_\pm$; see [9], § 12.4. We have
Although we know that $\mu_+(\boldsymbol{\theta})=\mu_-(\boldsymbol{\theta}) =: \mu(\boldsymbol{\theta})$, it is not so easy to check this relation using (19.10). In general, the operator (19.9) is not identically equal to zero, that is, Condition 10.2 is not satisfied. See [29], Example 16.2. The third-order operator $\widehat{N}_{Q,\pm}({\mathbf D})$ takes the form
We can apply Theorems 15.1, 15.5, and 15.12 and Corollaries 15.3, 15.9, and 15.15 to operators (19.11). Since now we have the case where $g^0=\underline{g}$, by Proposition 14.34 we can also apply the results ‘without smoothing’ (Theorems 15.22 and 15.25).
Applying Theorem 16.10 we obtain the following result.
Proposition 19.2. Suppose that ${u}_{\pm,\varepsilon}$ is the solution of problem (19.12) and ${u}_{0,\pm}$ is the solution of the homogenized problem (19.13). Let $v_{\pm,\varepsilon}$ be defined by (19.14), and let ${w}_{0,\pm}$ be the solution of problem (19.15).
If $\phi_\pm \in H^{s}(\mathbb{R}^2)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$$
\begin{equation*}
\|\omega_\mp^\varepsilon {u}_{\pm,\varepsilon}(\,{\cdot}\,, \tau)- \omega_{\mp,0} {u}_{0,\pm}(\,{\cdot}\,,\tau)\|_{L_2(\mathbb{R}^2)} \leqslant C (s) (1+|\tau|)^{s/3} \varepsilon^{s/3} \|{\phi}_\pm\|_{H^{s}(\mathbb{R}^2)}.
\end{equation*}
\notag
$$
If ${\phi}_\pm \in H^{s}(\mathbb{R}^2)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Applying Theorem 16.12, we obtain the following result.
Proposition 19.3. Suppose that $\tilde{u}_{\pm,\varepsilon}$ is the solution of problem (19.16) and ${u}_{0,\pm}$ is the solution of the homogenized problem (19.13). Let $\tilde{v}_{\pm,\varepsilon}$ be defined by (19.17), and let ${w}_{0,\pm}$ be the solution of problem (19.15).
If $\phi_\pm \in H^{s}(\mathbb{R}^2)$, where $1 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
19.3. The effective characteristics of the operator $\mathcal P$. Homogenization
The operator $\mathcal P$ is of the form (5.10), where $m=n=2$, $b({\mathbf D})=b_\times ({\mathbf D})$, $g({\mathbf x})=g_\times({\mathbf x})$, and $f({\mathbf x})=f_\times({\mathbf x})$. The role of the operator $\widehat{\mathcal A}$ for $\mathcal P$ is played by $\widehat{\mathcal A}_\times= b_\times ({\mathbf D}) g_\times({\mathbf x}) b_\times ({\mathbf D})$.
The role of the matrix-valued function $\Lambda({\mathbf x})$ for the operator $\mathcal P$ is played by $\Lambda_\times({\mathbf x})$, which is the $\Gamma$-periodic solution of the problem
where $g^0_\pm$ are defined by (19.6). The role of $Q({\mathbf x})$ for the operator $\mathcal P$ is played by the matrix $Q_\times({\mathbf x})= (f_\times({\mathbf x}))^{-2}=(g_\times({\mathbf x}))^{-1}$. Then
where $\Lambda_{Q,\pm}({\mathbf x})$ are the periodic functions defined in § 19.2.
Next, the operator $\widehat{N}_{Q,\times} (\boldsymbol{\theta})$, playing the role of $\widehat{N}_{Q}(\boldsymbol{\theta})$ for $\mathcal P$, takes the form
where the operators $\widehat{N}_{Q,\pm} (\boldsymbol{\theta})$ are defined by (19.9).
The first eigenvalue $\lambda(t,\boldsymbol{\theta})$ of the operator $\mathcal{P}(\mathbf k)$ is of multiplicity two for any $\mathbf k=t \boldsymbol{\theta}$, because the blocks $P_+(\mathbf k)$ and $P_-(\mathbf k)$ are unitarily equivalent. We have
where the coefficient $\gamma$ does not depend on $\boldsymbol{\theta}$ and is defined by (19.8), and the coefficient $\mu(\boldsymbol{\theta})$ is defined by (19.10). In general, $\mu(\boldsymbol{\theta})$ is not zero. The third-order operator $\widehat{N}_{Q,\times}({\mathbf D})$ is given by
We can apply Theorems 15.1, 15.5, and 15.12 and Corollaries 15.3, 15.9, and 15.15 to the operator (19.18). Since now the case where $g^0=\underline{g}$ is realized, by Proposition 14.34 we can also apply the results ‘without smoothing’ (Theorems 15.22 and 15.25).
Let ${\boldsymbol{\phi}}=\operatorname{col}\{\phi_{-},\phi_{+}\}$. It is clear that ${{\mathbf u}}_{\varepsilon}= \operatorname{col}\{u_{-,\varepsilon},u_{+,\varepsilon}\}$, where $u_{\pm,\varepsilon}$ are the solutions of problems (19.12).
Let $\mathbf{u}_{0}$ be the solution of the homogenized problem
Then ${\mathbf v}_\varepsilon= \operatorname{col}\{v_{-,\varepsilon},v_{+,\varepsilon}\}$, where $v_{\pm,\varepsilon}$ are defined by (19.14). Let ${\mathbf w}_{0}$ be the solution of the problem
Then ${{\mathbf w}}_0=\operatorname{col}\{w_{0,-},w_{0,+}\}$, where $w_{0,\pm}$ are the solutions of problems (19.15).
Applying Theorem 16.10 we obtain the following result.
Proposition 19.4. Suppose that ${{\mathbf u}}_{\varepsilon}$ is the solution of problem (19.19) and ${{\mathbf u}}_{0}$ is the solution of the homogenized problem (19.20). Let ${\mathbf v}_{\varepsilon}$ be defined by (19.21), and let ${{\mathbf w}}_{0}$ be the solution of problem (19.22).
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2;\mathbb{C}^2)$, where $0 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
$$
\begin{equation*}
\|f_\times^\varepsilon {{\mathbf u}}_{\varepsilon}(\,{\cdot}\,,\tau)- f_{\times,0}{{\mathbf u}}_{0}(\,{\cdot}\,,\tau)\|_{L_2 (\mathbb{R}^2)} \leqslant C (s) (1+|\tau|)^{s/3} \varepsilon^{s/3} \|\boldsymbol{\phi}\|_{H^{s}(\mathbb{R}^2)}.
\end{equation*}
\notag
$$
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2;\mathbb{C}^2)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2;\mathbb{C}^2)$, where $1 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $\varepsilon > 0$ we have
Let ${\boldsymbol{\phi}}=\operatorname{col}\{ \phi_{-},\phi_{+}\}$. Then $\tilde{{\mathbf u}}_{\varepsilon}=\operatorname{col} \{\tilde{u}_{-,\varepsilon},\tilde{u}_{+,\varepsilon}\}$, where $\tilde{u}_{\pm,\varepsilon}$ are the solutions of problems (19.16). Let ${\mathbf u}_0$ be the solution of the previous homogenized problem (19.20). We put
Then $\tilde{{\mathbf v}}_\varepsilon=\operatorname{col} \{\tilde{v}_{-,\varepsilon},\tilde{v}_{+,\varepsilon}\}$, where $\tilde{v}_{\pm,\varepsilon}$ are defined by (19.17).
Applying Theorem 16.12, we obtain the following result.
Proposition 19.5. Suppose that $\tilde{{\mathbf u}}_{\varepsilon}$ is the solution of problem (19.23), and let ${{\mathbf u}}_{0}$ be the solution of the homogenized problem (19.20). Let $\tilde{{\mathbf v}}_{\varepsilon}$ be defined by (19.24), and let ${{\mathbf w}}_{0}$ be the solution of problem (19.22).
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2; \mathbb{C}^2)$, where $1 \leqslant s \leqslant 3$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2;\mathbb{C}^2)$, where $3 \leqslant s \leqslant 6$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
If $\boldsymbol{\phi} \in H^{s}(\mathbb{R}^2;\mathbb{C}^2)$, where $2 \leqslant s \leqslant 4$, then for $\tau \in \mathbb{R}$ and $0 < \varepsilon \leqslant 1$ we have
A. Bensoussan, J.-L. Lions, and G. Papanicolaou, Asymptotic analysis for periodic structures, Stud. Math. Appl., 5, North-Holland Publishing Co., Amsterdam–New York, 1978, xxiv+700 pp.
2.
N. Bakhvalov and G. Panasenko, Homogenisation: averaging processes in periodic media. Mathematical problems in the mechanics of composite materials, Math. Appl. (Soviet Ser.), 36, Kluwer Acad. Publ., Dordrecht, 1989, xxxvi+366 pp.
3.
V. V. Jikov (Zhikov), S. M. Kozlov, and O. A. Oleinik, Homogenization of differential operators and integral functionals, Springer-Verlag, Berlin, 1994, xii+570 pp.
4.
E. V. Sevost'janova, “An asymptotic expansion of the solution of a second order elliptic equation with periodic rapidly oscillating coefficients”, Math. USSR-Sb., 43:2 (1982), 181–198
5.
V. V. Zhikov, “Spectral approach to asymptotic problems of diffusion”, Differ. Equ., 25:1 (1989), 33–39
6.
C. Conca, R. Orive, and M. Vanninathan, “Bloch approximation in homogenization and applications”, SIAM J. Math. Anal., 33:5 (2002), 1166–1198
7.
M. Sh. Birman and T. A. Suslina, “Second order periodic differential operators. Threshold properties and homogenization”, St. Petersburg Math. J., 15:5 (2004), 639–714
8.
M. Sh. Birman and T. A. Suslina, “Threshold approximations with corrector for the resolvent of a factorized selfadjoint operator family”, St. Petersburg Math. J., 17:5 (2006), 745–762
9.
M. Sh. Birman and T. A. Suslina, “Homogenization with corrector term for periodic elliptic differential operators”, St. Petersburg Math. J., 17:6 (2006), 897–973
10.
M. Sh. Birman and T. A. Suslina, “Homogenization with corrector for periodic differential operators. Approximation of solutions in the Sobolev class $H^1(\mathbb R^d)$”, St. Petersburg Math. J., 18:6 (2007), 857–955
11.
T. A. Suslina, “On homogenization of periodic parabolic systems”, Funct. Anal. Appl., 38:4 (2004), 309–312
12.
T. A. Suslina, “Homogenization of a periodic parabolic Cauchy problem”, Nonlinear equations and spectral theory, Amer. Math. Soc. Transl. Ser. 2, 220, Adv. Math. Sci., 59, Amer. Math. Soc., Providence, RI, 2007, 201–233
13.
E. S. Vasilevskaya, “A periodic parabolic Cauchy problem. Homogenization with corrector”, St. Petersburg Math. J., 21:1 (2010), 1–41
14.
T. Suslina, “Homogenization of a periodic parabolic Cauchy problem in the Sobolev space $H^1(\mathbb{R}^d)$”, Math. Model. Nat. Phenom., 5:4 (2010), 390–447
15.
E. S. Vasilevskaya and T. A. Suslina, “Threshold approximations for a factorized selfadjoint operator family with the first and second correctors taken into account”, St. Petersburg Math. J., 23:2 (2012), 275–308
16.
E. S. Vasilevskaya and T. A. Suslina, “Homogenization of parabolic and elliptic periodic operators in $L_2(\mathbb{R}^d)$ with the first and second correctors taken into account”, St. Petersburg Math. J., 24:2 (2013), 185–261
17.
T. A. Suslina, “Homogenization in the Sobolev class $H^1(\mathbb{R}^d)$ for second order periodic elliptic operators with the inclusion of first order terms”, St. Petersburg Math. J., 22:1 (2011), 81–162
18.
T. A. Suslina, “Homogenization of elliptic systems with periodic coefficients: operator error estimates in $L_2(\mathbb{R}^d)$ with corrector taken into account”, St. Petersburg Math. J., 26:4 (2015), 643–693
19.
Yu. M. Meshkova, “Homogenization of the Cauchy problem for parabolic systems with periodic coefficients”, St. Petersburg Math. J., 25:6 (2014), 981–1019
20.
Yu. M. Meshkova, “Homogenization of periodic parabolic systems in the $L_2(\mathbb{R}^d)$-norm with the corrector taken into account”, St. Petersburg Math. J., 31:4 (2020), 675–718
21.
V. V. Zhikov, “On operator estimates in homogenization theory”, Dokl. Math., 72:1 (2005), 534–538
22.
V. V. Zhikov and S. E. Pastukhova, “On operator estimates for some problems in homogenization theory”, Russ. J. Math. Phys., 12:4 (2005), 515–524
23.
V. V. Zhikov and S. E. Pastukhova, “Estimates of homogenization for a parabolic equation with periodic coefficients”, Russ. J. Math. Phys., 13:2 (2006), 224–237
24.
V. V. Zhikov and S. E. Pastukhova, “Operator estimates in homogenization theory”, Russian Math. Surveys, 71:3 (2016), 417–511
25.
M. Sh. Birman and T. A. Suslina, “Operator error estimates in the homogenization problem for nonstationary periodic equations”, St. Petersburg Math. J., 20:6 (2009), 873–928
26.
Yu. M. Meshkova, “On the homogenization of periodic hyperbolic systems”, Math. Notes, 105:6 (2019), 929–934
27.
Yu. M. Meshkova, “On operator error estimates for homogenization of hyperbolic systems with periodic coefficients”, J. Spectr. Theory, 11:2 (2021), 587–660
28.
Yu. M. Meshkova, Homogenization of periodic hyperbolic systems taking account of the corrector in the $L_2(\mathbb{R}^d)$-norm, Manuscript, 2018, 60 pp. (Russian)
29.
T. Suslina, “Spectral approach to homogenization of nonstationary Schrödinger-type equations”, J. Math. Anal. Appl., 446:2 (2017), 1466–1523
30.
M. A. Dorodnyi, “Operator error estimates for homogenization of the nonstationary Schrödinger-type equations: sharpness of the results”, Appl. Anal., 101:16 (2022), 5582–5614
31.
M. A. Dorodnyi and T. A. Suslina, “Spectral approach to homogenization of hyperbolic equations with periodic coefficients”, J. Differential Equations, 264:12 (2018), 7463–7522
32.
M. A. Dorodnyi and T. A. Suslina, “Homogenization of hyperbolic equations with periodic coefficients in ${\mathbb R}^d$: sharpness of the results”, St. Petersburg Math. J., 32:4 (2021), 605–703
33.
M. A. Dorodnyi, “Homogenization of periodic Schrödinger-type equations, with lower order terms”, St. Petersburg Math. J., 31:6 (2020), 1001–1054
34.
Yu. Meshkova, “Variations on the theme of the Trotter–Kato theorem for homogenization of periodic hyperbolic systems”, Russ. J. Math. Phys., 30:4 (2023), 561–598
35.
T. A. Suslina, “Homogenization of a stationary periodic Maxwell system”, St. Petersburg Math. J., 16:5 (2005), 863–922
36.
T. A. Suslina, “Homogenization with corrector for a stationary periodic Maxwell system”, St. Petersburg Math. J., 19:3 (2008), 455–494
37.
M. A. Dorodnyi and T. A. Suslina, “Homogenization of a non-stationary periodic Maxwell system in the case of constant permeability”, J. Differential Equations, 307 (2022), 348–388
38.
A. Piatnitski, V. Sloushch, T. Suslina, and E. Zhizhina, “On operator estimates in homogenization of nonlocal operators of convolution type”, J. Differential Equations, 352 (2023), 153–188
39.
Yu. Kondratiev, S. Molchanov, and E. Zhizhina, “On ground state of some non local Schrödinger operators”, Appl. Anal., 96:8 (2017), 1390–1400
40.
A. Piatnitski and E. Zhizhina, “Periodic homogenization of nonlocal operators with a convolution-type kernel”, SIAM J. Math. Anal., 49:1 (2017), 64–81
41.
D. I. Borisov, E. A. Zhizhina, and A. L. Piatnitskii, “Spectrum of a convolution operator with potential”, Russian Math. Surveys, 77:3 (2022), 546–548
42.
N. Veniaminov, “Homogenization of periodic differential operators of high order”, St. Petersburg Math. J., 22:5 (2011), 751–775
43.
A. A. Kukushkin and T. A. Suslina, “Homogenization of high order elliptic operators with periodic coefficients”, St. Petersburg Math. J., 28:1 (2017), 65–108
44.
V. A. Sloushch and T. A. Suslina, “Homogenization of the fourth-order elliptic operator with periodic coefficients with correctors taken into account”, Funct. Anal. Appl., 54:3 (2020), 224–228
45.
V. A. Sloushch and T. A. Suslina, “Threshold approximations for the resolvent of a polynomial nonnegative operator pencil”, St. Petersburg Math. J., 33:2 (2022), 355–385
46.
V. A. Sloushch and T. A. Suslina, “Operator estimates for homogenization of higher-order elliptic operators with periodic coefficients”, St. Petersburg Math. J., 35:2 (2024) (to appear)
47.
A. A. Miloslova and T. A. Suslina, “Homogenization of the higher-order parabolic equations with periodic coefficients”, J. Math. Sci. (N. Y.), 277 (2023), 959–1023
48.
T. A. Suslina, “Homogenization of the higher-order Schrödinger-type equations with periodic coefficients”, Partial differential equations, spectral theory, and mathematical physics, The Ari Laptev anniversary volume, EMS Ser. Congr. Rep., EMS Press, Berlin, 2021, 405–426
49.
T. A. Suslina, “Homogenization of the higher-order hyperbolic equations with periodic coefficients”, Lobachevskii J. Math., 42:14 (2021), 3518–3542
50.
S. E. Pastukhova, “Operator error estimates for homogenization of fourth order elliptic equations”, St. Petersburg Math. J., 28:2 (2017), 273–289
51.
S. E. Pastukhova, “Estimates in homogenization of higher-order elliptic operators”, Appl. Anal., 95:7 (2016), 1449–1466
52.
S. E. Pastukhova, “$L^2$-approximation of the resolvents in homogenization of higher order elliptic operators”, J. Math. Sci. (N. Y.), 251:6 (2020), 902–925
53.
S. E. Pastukhova, “Approximation of resolvents in homogenization of fourth-order elliptic operators”, Sb. Math., 212:1 (2021), 111–134
54.
S. E. Pastukhova, “Improved resolvent $L^2$-approximations in homogenization of fourth order operators”, St. Petersburg Math. J., 34:4 (2023), 611–634
55.
G. Griso, “Error estimate and unfolding for periodic homogenization”, Asymptot. Anal., 40:3-4 (2004), 269–286
56.
G. Griso, “Interior error estimate for periodic homogenization”, Anal. Appl. (Singap.), 4:1 (2006), 61–79
57.
C. E. Kenig, Fanghua Lin, and Zhongwei Shen, “Convergence rates in $L^2$ for elliptic homogenization problems”, Arch. Ration. Mech. Anal., 203:3 (2012), 1009–1036
58.
M. A. Pakhnin and T. A. Suslina, “Operator error estimates for homogenization of the elliptic Dirichlet problem in a bounded domain”, St. Petersburg Math. J., 24:6 (2013), 949–976
59.
T. A. Suslina, “Homogenization of the Dirichlet problem for elliptic systems: $L_2$-operator error estimates”, Mathematika, 59:2 (2013), 463–476
60.
T. Suslina, “Homogenization of the Neumann problem for elliptic systems with periodic coefficients”, SIAM J. Math. Anal., 45:6 (2013), 3453–3493
61.
T. A. Suslina, “Homogenization of elliptic operators with periodic coefficients in dependence of the spectral parameter”, St. Petersburg Math. J., 27:4 (2016), 651–708
62.
Yu. M. Meshkova and T. A. Suslina, Homogenization of the Dirichlet problem for elliptic systems: two-parametric error estimates, 2017, 45 pp., arXiv: 1702.00550
63.
Yu. M. Meshkova and T. A. Suslina, “Homogenization of the Dirichlet problem for elliptic and parabolic systems with periodic coefficients”, Funct. Anal. Appl., 51:3 (2017), 230–235
64.
Qiang Xu, “Convergence rates for general elliptic homogenization problems in Lipschitz domains”, SIAM J. Math. Anal., 48:6 (2016), 3742–3788
65.
Zhongwei Shen and Jinping Zhuge, “Convergence rates in periodic homogenization of systems of elasticity”, Proc. Amer. Math. Soc., 145:3 (2017), 1187–1202
66.
Zhongwei Shen, Periodic homogenization of elliptic systems, Oper. Theory Adv. Appl., 269, Adv. Partial Differ. Equ. (Basel), Birkhäuser/Springer, Cham, 2018, ix+291 pp.
67.
Yu. M. Meshkova and T. A. Suslina, “Homogenization of initial boundary value problems for parabolic systems with periodic coefficients”, Appl. Anal., 95:8 (2016), 1736–1775
68.
Yu. M. Meshkova and T. A. Suslina, “Homogenization of the first initial boundary-value problem for parabolic systems: operator error estimates”, St. Petersburg Math. J., 29:6 (2018), 935–978
69.
Jun Geng and Zhongwei Shen, “Convergence rates in parabolic homogenization with time-dependent periodic coefficients”, J. Funct. Anal., 272:5 (2017), 2092–2113
70.
T. A. Suslina, “Homogenization of a stationary periodic Maxwell system in a bounded domain in the case of constant magnetic permeability”, St. Petersburg Math. J., 30:3 (2019), 515–544
71.
T. A. Suslina, “Homogenization of the stationary Maxwell system with periodic coefficients in a bounded domain”, Arch. Ration. Mech. Anal., 234:2 (2019), 453–507
72.
T. A. Suslina, “Homogenization of the Dirichlet problem for higher-order elliptic equations with periodic coefficients”, St. Petersburg Math. J., 29:2 (2018), 325–362
73.
T. A. Suslina, “Homogenization of the Neumann problem for higher order elliptic equations with periodic coefficients”, Complex Var. Elliptic Equ., 63:7-8 (2018), 1185–1215
74.
T. A. Suslina, “Homogenization of higher-order parabolic systems in a bounded domain”, Appl. Anal., 98:1-2 (2019), 3–31
75.
Weisheng Niu, Zhongwei Shen, and Yao Xu, “Convergence rates and interior estimates in homogenization of higher order elliptic systems”, J. Funct. Anal., 274:8 (2018), 2356–2398
76.
Shu Gu, “Convergence rates in homogenization of Stokes systems”, J. Differential Equations, 260:7 (2016), 5796–5815
77.
D. I. Borisov, “Asymptotics for the solutions of elliptic systems with rapidly oscillating coefficients”, St. Petersburg Math. J., 20:2 (2009), 175–191
78.
S. E. Pastukhova and R. N. Tikhomirov, “Operator estimates in reiterated and locally periodic homogenization”, Dokl. Math., 76:1 (2007), 548–553
79.
S. E. Pastukhova and R. N. Tikhomirov, “Estimates of locally periodic and reiterated homogenization for parabolic equations”, Dokl. Math., 80:2 (2009), 674–678
80.
S. E. Pastukhova, “Approximation of the exponential of a diffusion operator with multiscale coefficients”, Funct. Anal. Appl., 48:3 (2014), 183–197
81.
S. E. Pastukhova, “On resolvent approximations of elliptic differential operators with locally periodic coefficients”, Lobachevskii J. Math., 41:5 (2020), 818–838
82.
N. N. Senik, “On homogenization for non-self-adjoint locally periodic elliptic operators”, Funct. Anal. Appl., 51:2 (2017), 152–156
83.
N. N. Senik, “On homogenization of locally periodic elliptic and parabolic operators”, Funct. Anal. Appl., 54:1 (2020), 68–72
84.
N. N. Senik, “Homogenization for locally periodic elliptic operators”, J. Math. Anal. Appl., 505:2 (2022), 125581, 24 pp.
85.
N. N. Senik, “Homogenization for locally periodic elliptic problems on a domain”, SIAM J. Math. Anal., 55:2 (2023), 849–881
86.
N. N. Senik, “On homogenization for piecewise locally periodic operators”, Russ. J. Math. Phys., 30:2 (2023), 270–274
87.
Weisheng Niu, Zhongwei Shen, and Yao Xu, “Quantitative estimates in reiterated homogenization”, J. Funct. Anal., 279:11 (2020), 108759, 39 pp.
88.
K. D. Cherednichenko and S. Cooper, “Resolvent estimates for high-contrast elliptic problems with periodic coefficients”, Arch. Ration. Mech. Anal., 219:3 (2016), 1061–1086
89.
K. D. Cherednichenko, Yu. Yu. Ershova, and A. V. Kiselev, “Effective behaviour of critical-contrast PDEs: micro-resonanses, frequency conversion, and time dispersive properties. I”, Comm. Math. Phys., 375:3 (2020), 1833–1884
90.
D. Borisov, G. Cardone, L. Faella, and C. Perugia, “Uniform resolvent convergence for strip with fast oscillating boundary”, J. Differential Equations, 255:12 (2013), 4378–4402
91.
D. Borisov, R. Bunoiu, and G. Cardone, “Waveguide with non-periodically alternating Dirichlet and Robin conditions: homogenization and asymptotics”, Z. Angew. Math. Phys., 64:3 (2013), 439–472
92.
D. I. Borisov and T. F. Sharapov, “On the resolvent of multidimensional operators with frequently alternating boundary conditions with the Robin homogenized condition”, J. Math. Sci. (N. Y.), 213:4 (2016), 461–503
93.
A. G. Chechkina, C. D'Apice, and U. De Maio, “Operator estimates for elliptic problem with rapidly alternating Steklov boundary condition”, J. Comput. Appl. Math., 376 (2020), 112802, 11 pp.
94.
V. V. Zhikov, “Spectral method in homogenization theory”, Proc. Steklov Inst. Math., 250 (2005), 85–94
95.
T. A. Suslina, “Spectral approach to homogenization of elliptic operators in a perforated space”, Rev. Math. Phys., 30:8 (2018), 1840016, 57 pp.
96.
K. Cherednichenko, P. Dondl, and F. Rösler, “Norm-resolvent convergence in perforated domains”, Asymptot. Anal., 110:3-4 (2018), 163–184
97.
S. E. Pastukhova, “Resolvent approximations in $L^2$-norm for elliptic operators acting in a perforated space”, J. Math. Sci. (N. Y.), 265:6 (2022), 1008–1026
98.
D. Borisov, G. Cardone, and T. Durante, “Homogenization and norm-resolvent convergence for elliptic operators in a strip perforated along a curve”, Proc. Roy. Soc. Edinburgh Sect. A, 146:6 (2016), 1115–1158
99.
D. I. Borisov and A. I. Mukhametrakhimova, “Uniform convergence and asymptotics for problems in domains finely perforated along a prescribed manifold
in the case of the homogenized Dirichlet condition”, Sb. Math., 212:8 (2021), 1068–1121
100.
D. I. Borisov, “Operator estimates for non-periodically perforated domains: disappearance of cavities”, Appl. Anal. (2023), Publ. online
101.
D. I. Borisov and J. Kříž, “Operator estimates for non-periodically perforated domains with Dirichlet and nonlinear Robin conditions: vanishing limit”, Anal. Math. Phys., 13:1 (2023), 5, 34 pp.
102.
A. Khrabustovskyi and O. Post, “Operator estimates for the crushed ice problem”, Asymptot. Anal., 110:3-4 (2018), 137–161
103.
C. Anné and O. Post, “Wildly perturbed manifolds: norm resolvent and spectral convergence”, J. Spectr. Theory, 11:1 (2021), 229–279
104.
A. Khrabustovskyi and M. Plum, “Operator estimates for homogenization of the Robin Laplacian in a perforated domain”, J. Differential Equations, 338 (2022), 474–517
105.
S. Brahim-Otsmane, G. A. Francfort, and F. Murat, “Correctors for the homogenization of the wave and heat equations”, J. Math. Pures Appl. (9), 71:3 (1992), 197–231
106.
G. Allaire and A. Piatnitski, “Homogenization of the Schrödinger equation and effective mass theorems”, Comm. Math. Phys., 258:1 (2005), 1–22
107.
S. Yu. Dobrokhotov, V. E. Nazaikinskii, and A. I. Shafarevich, “Efficient asymptotics of solutions to the Cauchy problem with localized initial data for linear systems of differential and pseudodifferential equations”, Russian Math. Surveys, 76:5 (2021), 745–819
108.
M. A. Dorodnyi and T. A. Suslina, “Homogenization of hyperbolic equations: operator estimates with correctors taken into account”, Funct. Anal. Appl., 57:4 (2023) (to appear)
109.
T. A. Suslina, “Threshold approximations of the exponential of a factorized operator family with correctors taken into account”, St. Petersburg Math. J., 35:3 (2024) (to appear)
110.
T. Kato, Perturbation theory for linear operators, Grundlehren Math. Wiss., 132, Springer-Verlag New York, Inc., New York, 1966, xix+592 pp.
111.
V. G. Maz'ya and T. O. Shaposhnikova, Theory of multipliers in spaces of differentiable functions, Monogr. Stud. Math., 23, Pitman, Boston, MA, 1985, xiii+344 pp.
112.
O. A. Ladyzhenskaya and N. N. Ural'tseva, Linear and quasilinear elliptic equations, Academic Press, New York–London, 1968, xviii+495 pp.
113.
R. G. Shterenberg, “On the structure of the lower edge of the spectrum of the periodic magnetic Schrödinger operator with small magnetic potential”, St. Petersburg Math. J., 17:5 (2006), 865–873
Citation:
T. A. Suslina, “Operator-theoretic approach to the homogenization of Schrödinger-type equations with periodic coefficients”, Russian Math. Surveys, 78:6 (2023), 1023–1154