\input amstex \documentstyle{amsppt} \magnification=1200 \baselineskip=15 pt \TagsOnRight \topmatter \title Absolute Summability of the Trace Relation \\ for Certain Schr\"odinger Operators \endtitle \rightheadtext{Absolute Summability of the Trace Relation} \author F. Gesztesy$^{1}$, H. Holden$^{2}$, and B. Simon$^{3}$ \endauthor \leftheadtext{F. Gesztesy, H. Holden, and B. Simon} \thanks $^{1}$ Department of Mathematics, University of Missouri, Columbia, MO 65211. E-mail:mathfg\@mizzou1.\linebreak missouri.edu \endthanks \thanks $^{2}$ Department of Mathematical Sciences, The Norwegian Institute of Technology, University of Trondheim, N-7034 Trondheim, Norway. E-mail: holden\@imf.unit.no \endthanks \thanks $^{3}$ Division of Physics, Mathematics and Astronomy, California Institute of Technology, 253-37,\linebreak Pasadena, CA 91125. This material is based upon work supported by the National Science Foundation under Grant No. DMS-9101715. The Government has certain rights in this material. \endthanks \thanks To be submitted to {\it Commun.~Math.~Phys.} \endthanks \abstract A recently established general trace formula for one-dimensional Schr\"odinger operators is systematically studied in the context of short-range potentials, potentials which approach different spatial asymptotes sufficiently fast, and appropriate impurity (defect) interactions in one-dimensional solids. We prove the absolute summability of the trace formula and establish its connections with scattering quantities, such as reflection coefficients, in each case. \endabstract \endtopmatter \document \bigpagebreak \flushpar {\bf \S 1. Introduction} This paper is the third in a series on a general trace formula and its ramifications in (inverse) spectral theory for one-dimensional Schr\"odinger operators started in [16] and continued in [19]. The main theme in [16] concentrates around a general trace formula for self-adjoint Schr\"odinger operators $H$ in $L^{2}(\Bbb R)$ of the type $$ H=-\frac{d^2}{dx^2}+V, \tag 1.1 $$ where we assume that the real-valued potential $V(x)$ is continuous and bounded from below. In order to gain some information on $V(y)$, we shall compare $H$ with the associated self-ajoint Dirichlet operator $H^{D}_{y}$ obtained from $H$ by imposing an additional Dirichlet boundary condition $\lim\limits_{\epsilon\downarrow 0}\psi(y\pm\epsilon)=0$ at the point $y\in\Bbb R$. Since the resolvent difference $[(H^{D}_{y}-z)^ {-1}-(H-z)^{-1}]$ is rank one (cf.~(2.37)), Krein's spectral shift function $\xi (\lambda, y)$ [25], [33] for the pair $(H^{D}_{y}, H)$ exists for all $y\in\Bbb R$ and a.e.~$\lambda\in\Bbb R$ (with respect to Lebesgue measure) and one obtains for all $y\in\Bbb R$ $$ \text{Tr}[f(H^{D}_{y})-f(H)]=\int\limits_{\Bbb R} d\lambda f'(\lambda)\xi(\lambda, y), \tag 1.2 $$ $$ 0\leq\xi(\lambda, y)\leq 1\quad\text{a.e.~$\lambda\in\Bbb R$}, \tag 1.3 $$ $$ \xi (\lambda, y)=0,\quad\lambda<\inf\,\sigma(H) \tag 1.4 $$ for any $f\in C^{2}(\Bbb R)$ with $(1+\lambda^{2})f^{(j)}\in L^{2}((0,\infty))$, $j=1, 2$ and $f(\lambda)=(\lambda - z)^{-1}$, $z\in\Bbb C\backslash [\inf\, \sigma(H), \infty)$. (Here $\sigma(\cdot)$ denotes the spectrum.) A closer look at the rank-one resolvent difference of $H^{D}_{y}$ and $H$ reveals the additional result that for each $y\in\Bbb R$ and a.e.~$\lambda\in\Bbb R$, $$ \xi (\lambda, y)=\lim_{\epsilon\downarrow 0} \pi^{-1}\text{Im} \{\ln [G(\lambda +i\epsilon , y, y)]\}, \tag 1.5 $$ where $G(z, x, x')$ denotes the Green's function of $H$ (i.e., the integral kernel of $(H-z)^{-1}$). The main results proven in [16], [19] then revolve around the following general trace formula. \proclaim{Theorem 1.1} \rom{[16,19]} Let $V$ be a measurable function on $\Bbb R$ satisfying \roster \item"\rom{(i)}" $\sup\limits_{n\in\Bbb N}\int\limits^{n+1}_{n} dx |V_{-}(x)|<\infty$, \item"\rom{(ii)}" $\int\limits^{n+1}_{n} dx |V_{+}(x)|<\infty$ for all $n\in\Bbb N$, \endroster where $V_{\pm}(x)=[|V(x)|\pm V(x)]/2$ and suppose $E_{o}\leq\inf\,\sigma(H)$. If $x$ is a point of Lebesgue continuity for $V$, then $$ V(x)=E_{o}+\lim_{\epsilon\downarrow 0} \int\limits^{\infty}_{E_{o}} d\lambda e^{-\epsilon\lambda}[1-2\xi (\lambda, x)]. \tag 1.6 $$ \endproclaim The proof of (1.6) combines (1.2) for $f(\lambda)=e^{-\epsilon\lambda}$, $\epsilon >0$ with path integral arguments to control the trace of the heat kernel difference as $\epsilon\downarrow 0$. In the particularly simple case $V(x)\equiv 0$, $G(\lambda, x, x)=i/\lambda^{1/2}$, Im$(\lambda^{1/2})\geq 0$ for $\lambda\geq 0$ and hence $\xi(\lambda, x)=\cases 1/2, \lambda >0 \\0, \lambda <0\endcases$. Further explicit examples can be found in Remark 2.5 in the context of reflectionless ($N$-soliton) potentials and in (4.18)--(4.20) in connection with periodic potentials. In fact, historically, after the pioneering work by Gel'fand and Levitan [12] on regularized traces for Schr\"odinger operators on a compact interval, the trace formula (4.19) for periodic (and certain classes of almost periodic) potentials was one of the two previously systematically studied trace formulae of the type (1.6) for Schr\"odinger operators on the whole real line (see, e.g., [8], [11], [22], [30], [34] and more recently [5], [24], [27], [28]). The other case studied in detail by Deift and Trubowitz [7] in 1979 was concerned with short-range potentials $V(x)$ decaying sufficiently fast as $|x|\to\infty$ under the assumption that $H=-\frac {d^2}{dx^2}+V$ has no eigenvalues. They proved that $$ V(x)=\frac{2i}{\pi}\int\limits^{\infty}_{-\infty}dk\,k\,\ln\biggl[ 1+R(k)\frac{f_{+}(k, x)}{f_{-}(k, x)}\biggr] \tag 1.7 $$ (where $f_{\pm}(k, x)$ are the Jost functions at energy $E=k^2$ and $R(k)$ is a reflection coefficient) which is an analog of (1.6). In the special case of positive $C^\infty$-potentials of compact support, a trace formula of the type $$ V(x)=\int\limits^{\infty}_{0} d\lambda[1-2\xi(\lambda, x)], \quad x\in\Bbb R \tag 1.8 $$ has recently been established by Venakides [35]. However, the equivalence of (1.7) and (1.8) was not established in [35]. Moreover, $\xi(\lambda, x)$ was not identified as Krein's spectral shift function for the pair $(H^{D}_{x}, H)$ and also the connection (1.5) between $\xi(\lambda, x)$ and the Green's function of $H$ was not made. The analog of the trace formula (1.6) and the associated formalism for second-order finite-difference (Jacobi) operators, a summability result for operators $H$ with purely discrete spectrum along with a powerful new characterization of the absolutely continuous spectrum $\sigma_{\text{ac}}(H)$ of $H$ as $$ \sigma_{\text{ac}}(H)=\overline{\{\lambda\in\Bbb R\mid 0<\xi(\lambda, x)<1\}}^{\text{ ess}} \tag 1.9 $$ for each fixed $x\in\Bbb R$ and some of its applications to the almost Mathieu equation or Harper's model (here $^{\text{---ess}}$ denotes the essential closure) are presented in the first paper [16] of our series. The case of general self-adjoint boundary conditions of the type $\psi'(x)+\beta\psi(x)=0$, $\beta\in\Bbb R\cup\{\infty\}$ together with trace formulas for all higher-order KdV invariants (expressed as differential polynomials in $V$) are studied in detail in the second paper [19] of our series. In the present third paper of our series, we shall give a systematic study of short-range perturbations in which the regularization (Abelian limit) $\epsilon\downarrow 0$ in (1.6) can be removed and prove the absolute summability of the trace formula (1.6). Specifically, we shall study the following three situations: \roster \item"\rom{(i)}" Sufficiently short-range potentials with certain regularity properties (typically, $V\in H^{2,1}(\Bbb R)$, see (2.1)) in \S 2. \item"\rom{(ii)}" Potentials which tend to different asymptotes as $x\to\pm\infty$ sufficiently fast in \S 3. \item"\rom{(iii)}" Impurity (defect) scattering in connection with potentials of the type $V=V^{o}+W$, where $V^{o}(x+a)=V^{o}(x)$ for some $a>0$ represents the periodic background and the short-range perturbation $W$ models impurities (defects) in a one-dimensional crystal, are treated in \S 4. \endroster In each of these three situations, we establish the connection between $\xi(\lambda, x)$ and appropriate scattering quantities, such as reflection coefficients and Jost functions, and prove the absolute summability of the trace formula (1.6), $$ \int\limits^{\infty}_{R}d\lambda |1-2\xi(\lambda, x)|<\infty, \quad R\in\Bbb R \tag 1.10 $$ removing the Abelian limit $\epsilon\downarrow 0$ in (1.6). It should be pointed out at this occasion that the Abelian limit $\epsilon\downarrow 0$ in (1.6) cannot be removed in general if $V(x)\to\infty$ as $x\to\infty$ or $x\to -\infty$ irrespective of the regularity properties of $V(x)$. This is particularly plain in the case where $V(x)\to\infty$ as $x\to\pm\infty$ since then for each $x\in\Bbb R$, $|1-2\xi(\lambda, x)|=1$ for a.e.~$\lambda\in\Bbb R$. But even if $V(x)$ tends to a constant sufficiently fast as $x\to - \infty$ and $V(x)\operatornamewithlimits{\longrightarrow} \limits_{x\to\infty}\infty$, explicit examples (such as, e.g., $V(x)=e^x$) in [26] show that $[1-2\xi(., x)]\notin L^{1} ((R, \infty); d\lambda), R, x\in\Bbb R$. In these situations, the Abelian limit $\epsilon\downarrow 0$ in (1.6) represents a genuine summability method. The fourth paper [17] in our series is devoted to various multidimensional trace formulas in terms of heat kernel asymptotics. A brief announcement of our results appeared in [18], expository accounts of this circle of ideas can be found in [14], [33]. Papers exploring several new solutions of inverse spectral problems are in preparation. \bigpagebreak \flushpar {\bf \S 2. Short-range Potentials} In this section we illustrate the trace formula (1.6) in the particular case of short-range potentials satisfying $$ V\in H^{2,1}(\Bbb R), \quad\text{$V$ real-valued.} \tag 2.1 $$ Here $H^{m, p}(\Bbb R)$, $m, p\in\Bbb N$ denotes the usual Sobolev space whose elements have up to $m$ distributional derivatives in $L^{p}(\Bbb R)$. The regularity condition on $V$ in (2.1) is essential in connection with our main result in Theorem 2.3, the removal of a regularization procedure (Abelian limit) in our trace formula (1.6) (cf. (2.39)--(2.41)). The associated self-adjoint Schr\"odinger operator $H$ in $L^{2}(\Bbb R)$ is then defined by $$ H=-\frac {d^2}{dx^2}+V, \quad \Cal D(H)=H^{2, 2}(\Bbb R) \tag 2.2 $$ and the spectrum $\sigma (H)$ of $H$ is of the type $$ \sigma (H)=\sigma_{d}(H)\cup [0, \infty), \quad \sigma_{\text{ess}}(H)=[0, \infty). \tag 2.3 $$ Here $\sigma_{\text{ess}}(H)$ is the essential spectrum of $H$ and the discrete spectrum $\sigma_{d}(H)$ of $H$ is a bounded subset of $(-\infty, 0)$ which may be empty, finite, or countably infinite. We denote the latter by $$ \sigma_{d}(H)=\{e_{j}\}_{j\in J} \ ,\quad e_{j} 0$ in $\Bbb C^2$ associated with the pair $(H, H_o)$ then explicitly reads in terms of transmission and reflection coefficients from left and right incidence $$ S(\lambda)=\pmatrix T(\lambda) & R^{r}(\lambda) \\ R^{\ell}(\lambda) & T(\lambda) \endpmatrix , \quad \lambda >0, \tag 2.14 $$ where $$ T(\lambda)=\frac {2i\lambda^{1/2}}{W(f_{-}(\lambda), f_{+}(\lambda))} = \left[ 1-\frac {1}{2i\lambda^{1/2}}\int\limits_{\Bbb R}dxV(x)e^{\pm i\lambda^{1/2}x} f_{\mp}(\lambda, x)\right]^{-1}, \tag 2.15 $$ $$ R^{\ell}(\lambda)=-\frac {W(\overline{f_{-}(\lambda)}, f_{+}(\lambda))}{W(f_{-}(\lambda), f_{+}(\lambda))} =\frac {T(\lambda)}{2i\lambda^{1/2}} \int\limits_{\Bbb R} dxV(x)e^{i\lambda^{1/2}x}f_{+}(\lambda, x), \tag 2.16 $$ $$ R^{r}(\lambda)=-\frac {W(f_{-}(\lambda), \overline{f_{+}(\lambda)})} {W(f_{-}(\lambda), f_{+}(\lambda))} = \frac {T(\lambda)}{2i\lambda^{1/2}} \int\limits_{\Bbb R} dxV(x) e^{-i\lambda^{1/2}x}f_{-}(\lambda, x), \tag 2.17 $$ where $$ f_{\pm}(\lambda, x)=\lim_{\epsilon\downarrow 0} f_{\pm}(\lambda +i\epsilon, x), \quad \lambda > 0 \tag 2.18 $$ and $W(f, g)(x)=f(x)g'(x)-f'(x)g(x)$ denotes the Wronskian of $f$ and $g$. In addition, we recall that $$ T(\lambda)f_{\pm}(\lambda, x)=R\Sp \ell \\ r\endSp(\lambda)f_{\mp} (\lambda, x) + \overline{f_{\mp}(\lambda, x)}, \quad \lambda >0 \tag 2.19 $$ and that the Green's function of $H$ (the integral kernel of $(H-z)^ {-1}$) is given by $$ G(z, x, x')=\frac {f_{+}(z, x)f_{-}(z, x')}{W(f_{+}(z), f_{-}(z))}, \quad x'\leq x, z\in\Bbb C\backslash\{0\}, \text{Im}(z^{1/2})\geq 0. \tag 2.20 $$ Specializing to $x=x', G(z, x, x)$ is well known to be a Herglotz function in $z\in\Bbb C_{+}=\mathbreak\{z\in\Bbb C\mid\text{Im}(z)>0\}$ for all $x\in\Bbb R$, that is, $G(z, x, x)$ is analytic in $\Bbb C_{+}$ and $$ \text{Im}[G(z, x, x)]>0,\quad\overline{G(z, x, x)}=G(\bar{z}, x, x), \quad z\in\Bbb C_{+}, x\in\Bbb R. \tag 2.21 $$ As a consequence (see, e.g., [1] or, for a different approach, [16], [33]) $G(z, x, x)$ admits the exponential representation $$ G(z, x, x)=\exp\left[c+\int\limits_{\Bbb R}\left[\frac{1}{\lambda-z} -\frac{\lambda}{1+\lambda^2}\right]\, \xi(\lambda, x)d\lambda\right], \quad z\in\Bbb C_{+}, x\in\Bbb R, \tag 2.22 $$ where $$ c\in\Bbb R,\quad 0\leq\xi(\lambda, x)\leq 1 \quad\text{for a.e.~$\lambda\in\Bbb R$}. \tag 2.23 $$ Fatou's lemma then implies that $$ \xi(\lambda, x)=\lim_{\epsilon\downarrow 0}\pi^{-1} \text{Im}\{\ln[G(\lambda +i\epsilon, x, x)]\} \tag 2.24 $$ exists for all $x\in\Bbb R$ and a.e.~$\lambda\in\Bbb R$. The normalization $$ \xi(\lambda, x)=0,\quad\lambda 0, \quad\lambda0. \tag 2.30 $$ In particular, $\xi(\lambda, x)$ is continuous for $\lambda>0$. Moreover, $$ |1-2\xi(\lambda, x)|\leq |R^{r(\ell)}(\lambda)|, \quad \lambda >0 \tag 2.31 $$ and $$ [1-2\xi(\lambda, x)]\operatornamewithlimits{=}_{\lambda\to\infty} o(\lambda^{-3/2}) \tag 2.32 $$ uniformly with respect to $x\in\Bbb R$. In addition, $$\align &\xi(\lambda, x)=\cases 0, &\text{\rom{if} $\lambda 0 \endaligned \tag 2.35 $$ which in turn is implied by (2.15), (2.18)--(2.20). Continuity of $\xi(\lambda, x)$ for $\lambda >0$ follows from the fact that $G(\lambda +i0, x, x)$ is continuous and zero-free for $\lambda\in (0, \infty)$. Inequality (2.31) follows from (2.30) and an elementary geometrical argument. (In fact, $|\arg(1+Re^{i\varphi})|\leq \arcsin (|R|)\leq\pi/2$ for $|R|\leq 1, \varphi\in\Bbb R$ and $\sin(x)\geq 2x/\pi$ for $0\leq x\leq\pi/2$ imply $|\arg(1+Re^{i\varphi})|\leq\pi |R|/2$.) Relation (2.32) is then implied by (2.31) and $$ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} o(\lambda^{-3/2}) \tag 2.36 $$ which is a consequence of (2.15)--(2.17) and two integrations by parts applying the \linebreak Riemann-Lebesgue lemma. Relation (2.33) directly follows from (2.24) and the fact that $G(z, x, x)$ is real-valued for $z<0$ with zeros precisely at the Dirichlet eigenvalues $\mu_{j}(x)$ of $H^{D}_{x}$ since $$ (H^{D}_{x}-z)^{-1}=(H-z)^{-1}-G(z, x, x)^{-1}(\overline{G(z, x, .)}, .) G(z, ., x), \, z\in\Bbb C\backslash\{\sigma(H^{D}_{x})\cup\sigma(H)\}. \tag 2.37 $$ Here $(., .)$ denotes the scalar product in $L^{2}(\Bbb R)$. Relation (2.34) then follows from (2.33) by a continuity argument. \qed \enddemo \remark{Remark \rom{2.2}} (i) Inequality (2.31) holds for any real-valued potential satisfying $V\in L^{1}(\Bbb R)$ and hence $$ [1-2\xi(., x)]\in L^{1}((0, \infty); d\lambda) \ \text{if} \ R^{r(\ell)} \in L^{1}((0, \infty); d\lambda). \tag 2.38 $$ In particular, $R^{r(\ell)}\in L^{1}((0, \infty); d\lambda)$ together with (2.31) are all that's needed to remove the Abelian limit in (2.27) (see Theorem 2.3). We also note that (2.36) (and hence (2.32)) holds if $V'$ is merely piecewise absolutely continuous admitting finitely-many jump discontinuities (i.e., there exists a finite partition of $\Bbb R, -\infty =x_{0}0$. In this case $\xi(\lambda, x)$ is given by (2.33), (2.34) for $\lambda <0$ and by $$ \xi(\lambda, x)=\frac12 , \quad \lambda >0 \tag 2.42 $$ on the (interior of the) absolutely continuous spectrum of $H$. This applies, in particular, to all $N$-soliton potentials (including $V\equiv 0$) and to a class of $\infty$-soliton potentials (having infinitely-many negative eigenvalues accumulating at zero) introduced in [20], [21]. \endremark We conclude this section with a few remarks on the low-energy behavior of $\xi(\lambda, x)$ as $\lambda\downarrow 0$. Assuming, in addition to (2.1), that $V$ satisfies $$ V\in L^{1}(\Bbb R; (1+x^{2})dx), \tag 2.43 $$ we need to consider the following case distinctions: \example{Case I} $W(f_{-}(0), f_{+}(0))\neq 0$ and $f_{-}(0, x) f_{+}(0, x)\neq 0.$ \endexample \flushpar (The first requirement can be expressed as $\int\limits_{\Bbb R} dx V(x)f_{\pm}(0, x)\neq 0$ and is equivalent to the fact that $H$ has no threshold resonance; see, e.g. [2], [3]. The second requirement says $\lim\limits_{x'\to x}\mu_{N+1}(x')\neq 0$.) Then $$ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow 0} -1+O(\lambda^{1/2}), \tag 2.44 $$ $$T(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow 0} \frac{-2i\lambda^{1/2}}{\int\limits_{\Bbb R}dx'V(x')f_{\mp}(0, x')} [1+O(\lambda^{1/2})], \tag 2.45 $$ $$ G(\lambda +i0, x, x)\operatornamewithlimits{=}_{\lambda\downarrow 0} \frac{-f_{+}(0, x)f_{-}(0, x)}{\int\limits_{\Bbb R}dx'V(x')f_{\pm}(0, x')} +O(\lambda^{1/2}) \tag 2.46 $$ and hence $$ \xi(\lambda, x)=\pi^{-1}\arg[G(\lambda+i0, x, x)] \operatornamewithlimits{=}_{\lambda\downarrow 0} O(\lambda^{1/2}) \ \text{in case I} \tag 2.47 $$ since $f_{\pm}(0, x)$ are real-valued. \example{Case II} $W(f_{-}(0), f_{+}(0))=0$ and $f_{-}(0, x)f_{+}(0, x)\neq 0$. \endexample \flushpar (The first requirement can be written as $\int\limits_{\Bbb R} dxV(x)f_{\pm}(0, x)=0$ and is equivalent to the fact that $H$ has a threshold resonance, see, e.g., [2], [3].) Then [2], [3] $$ R\Sp r\\ \ell\endSp(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow 0} \mp \frac{2c_{1}\overline{c_{2}}}{|c_{1}|^{2}+|c_{2}|^{2}} +O(\lambda^{1/2}), \tag 2.48 $$ $$ T(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow 0} \frac{|c_{1}|^{2}-|c_{2}|^{2}}{|c_{1}|^{2}+|c_{2}|^{2}} +O(\lambda^{1/2}), \tag 2.49 $$ with $|c_{1}|\neq |c_{2}|$ and hence $R^{r(\ell)}(0)\neq -1$. Thus $$ G(\lambda+i0, x, x)\operatornamewithlimits{=}_{\lambda\downarrow 0} (i/2\lambda^{1/2})|f_{\pm}(0, x)|^{2}[1+R\Sp r\\ \ell\endSp(0)] +O(1) \tag 2.50 $$ and hence $$ \xi(\lambda, x)=\pi^{-1}\arg[G(\lambda+i0, x, x)] \operatornamewithlimits{=}_{\lambda\downarrow 0} \frac12 +O(\lambda^{1/2}) \ \text {in case II}. \tag 2.51 $$ We emphasize that $V\equiv 0$ and more generally, all $N$-soliton potentials $V_N$ mentioned in Remark 2.5 have a zero-energy resonance and hence belong to case II. The case where $f_{-}(0, x)f_{+}(0, x)=0$ can be dealt with analogously, but requires higher-order computations. \bigpagebreak \flushpar {\bf \S 3. Cascades} As the the title of this section suggests, we shall now indicate how to extend the results of \S 2 to potentials with non-trivial spatial asymptotics. More precisely, we shall assume that $V$ satisfies $$\gathered V, V'\in AC_{\text{loc}}(\Bbb R), \ V \text{real-valued},\ V', V''\in L^{1}(\Bbb R),\\ \int\limits^{0}_{-\infty}dx|V(x)|+ \int\limits^{\infty}_{0}dx|V(x)-V_{+}|<\infty \ \text{for some $V_{+}>0$} \endgathered \tag 3.1 $$ throughout the major part of this section. (By reflection, $x\to -x$, it suffices to consider $V_{+}>0$.) Since most of the details will be similar to those in the previous section, we shall mostly refer to \S2 for notations and basic facts and dwell only on situations markedly different in the present context of (3.1). Introducing $H$, $H^{D}_{y}$, $J$, $J_{+}$, etc.~as in \S 2, the absolutely continuous spectrum of $H$ and $H^{D}_{y}$ now equals $[0, \infty)$ with uniform spectral multiplicity one on $(0, V_{+})$ and two on $(V_{+}, \infty)$. While $H$ has no embedded eigenvalues in $(0, \infty)$, $H^{D}_{y}$ may have (countably infinitely-many) eigenvalues in $[0, V_{+}]$ as briefly discussed in Remark 3.4. Concerning the stationary scattering theory for $H$, one has to replace the Jost solutions (2.12) by $$\gathered f_{\pm}(z, x)=e^{\pm ik_{\pm}x}-\int\limits^{\pm\infty}_{x} dx'k^{-1}_{\pm}\sin[k_{\pm}(x-x')][V(x')-V_{\pm}]f_{\pm}(z, x'), \\ k_{+}(z)=(z-V_{+})^{1/2}, \ k_{-}(z)=z^{1/2}, \ V_{-}=0, \ \text{Im} [k_{\pm}(z)]\geq 0, \ z\in\Bbb C\backslash\{0, V_{+}\}, \ x\in\Bbb R \endgathered \tag 3.2 $$ to obtain $$ Hf_{\pm}(z, x)=zf_{\pm}(z, x), \quad z\in\Bbb C\backslash\{0, V_{+}\} \tag 3.3 $$ in the distributional sense. The unitary scattering matrix $S(\lambda)$, $\lambda >0$ in $\Bbb C$ resp. $\Bbb C^2$ now reads as follows: $$ S(\lambda)=R^{\ell}(\lambda)=-\frac{\overline{W(f_{-}(\lambda), f_{+}(\lambda))}}{W(f_{-}(\lambda), f_{+}(\lambda))}, \quad 0<\lambda V_{+}, \tag 3.5 $$ where $$ T(\lambda)=\frac{2i[k_{+}(\lambda)k_{-}(\lambda)]^{1/2}} {W(f_{-}(\lambda), f_{+}(\lambda))}, \tag 3.6 $$ $$ R^{\ell}(\lambda)=-\frac{W(\overline{f_{-}(\lambda)}, f_{+}(\lambda))} {W(f_{-}(\lambda), f_{+}(\lambda))}, \tag 3.7 $$ $$ R^{r}(\lambda)=-\frac{W(f_{-}(\lambda), \overline{f_{+}(\lambda)})} {W(f_{-}(\lambda), f_{+}(\lambda))}. \tag 3.8 $$ Here $f_{\pm}(\lambda, x)=\lim\limits_{\epsilon\downarrow 0} f_{\pm}(\lambda+i\epsilon, x)$ and the Green's function $G(z, x, x)$ of $H$ now satisfies $$\gathered G(\lambda +i0, x, x)=\frac{f_{+}(\lambda, x)f_{-}(\lambda, x)} {W(f_{+}(\lambda), f_{-}(\lambda))} \\ =[i/2k_{\pm}(\lambda)]|f_{\pm}(\lambda, x)|^{2}\left[1+R\Sp r\\ \ell\endSp (\lambda)\frac{f_{\pm}(\lambda, x)^{2}}{|f_{\pm}(\lambda, x)|^{2}}\right], \quad \lambda > \cases V_{+} \\ 0 \endcases , \quad x\in\Bbb R. \endgathered \tag 3.9 $$ The above expression for $G(\lambda+i0, x, x)$ involving $R^{r}(\lambda)$ (as opposed to that involving $R^{\ell}(\lambda)$) appears to be singular at $\lambda = V_+$ since $k_{+}(\lambda)^{- 1}=(\lambda-V_{+})^{-1/2}$ (whereas $k_{-}(V_{+})^{-1}=V^{-1/2}_{+}$). However, this apparent contradiction is easily resolved by observing that $R^{r}(\lambda)\operatornamewithlimits{=}\limits_{\lambda\downarrow V_{+}}-1+o(1)$ (see also (3.33) for more details). \remark{Remark \rom{3.1}} Scattering theory for potentials with different spatial asymptotics has been studied in detail, for example, in [4], [6], [13], and we have freely used these results in (3.2)--(3.9). That $S(\lambda)$ for $0,\lambda \cases V_{+} \\ 0 \endcases \tag 3.10 $$ and $\xi(\lambda, x)$ is continuous in $\lambda >V_{+}$. \endproclaim \demo{Proof} Except for the analog of (2.32), which follows again from $$ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} o(\lambda^{-3/2}), \tag 3.11 $$ everything else is proven as in Lemma 2.1. The actual proof of (3.11), however, is now more cumbersome since no simple formulas such as the right-hand sides in (2.15)--(2.17) appear to be available in the present case. Hence, we briefly sketch a different (though straightforward) approach to (3.11) (following Lemma 2.3 in [13]). >From the outset it is readily verified that $$\gather T(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} 1+O(\lambda^{- 1/2}), \tag 3.12 \\ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} O(\lambda^{-1/2}). \tag 3.13 \endgather $$ In order to improve on (3.13), using the additional smoothness conditions on $V$, we explicitly compute the Wronskian of $\overline{f_{-}(\lambda, x)}$ and $f_{+}(\lambda, x)$ (for simplicity, at $x=0$). $$\gather W(\overline{f_{-}(\lambda)}, f_{+}(\lambda))(0)=i(k_{+}-k_{-})- \int\limits^{\infty}_{0} dx \cos(k_{+}x)[V(x)-V_{+}] f_{+}(\lambda, x) \\ -\int\limits^{0}_{-\infty}dx \cos(k_{-}x)V(x)\overline{f_{-}(\lambda, x)} -(ik_{+}/k_{-})\int\limits^{0}_{-\infty}dx\sin(k_{-}x)V(x) \overline{f_{-}(\lambda, x)} \\ -(ik_{-}/k_{+})\int\limits^{\infty}_{0}dx\sin(k_{+}x)[V(x)- V_{+}]f_{+}(\lambda, x) \\ +k^{-1}_{-} \int\limits^{0}_{-\infty}dx\sin(k_{-}x/V(x) \overline{f_{-}(\lambda, x)} \int\limits^{\infty}_{0} dx'\cos(k_{+}x')[V(x')-V_{+}] f_{+}(\lambda, x')\\ -k^{-1}_{+} \int\limits^{0}_{-\infty} dx\cos(k_{-}x)V(x) \overline{f_{-}(\lambda, x)} \int\limits^{\infty}_{0} dx' \sin(k_{+}x') [V(x')-V_{+}] f_{+}(\lambda, x'), \quad \lambda > V_{+}. \tag 3.14 \endgather $$ Next, one observes $$ i[k_{+}(\lambda)-k_{-}(\lambda)]\operatornamewithlimits{=} _{\lambda\to\infty}(-i/2\lambda^{1/2})V_{+}+O(\lambda^{-3/2}), \tag 3.15 $$ $$[ik_{\pm}(\lambda)/k_{\mp}(\lambda)]\operatornamewithlimits{=} _{\lambda\to\infty} i\mp (i/2\lambda)V_{+}+O(\lambda^{-2}), \tag 3.16 $$ $$ k_{\pm}(\lambda)^{-1}\operatornamewithlimits{=}_{\lambda\to\infty} \lambda^{-1/2}+O(\lambda^{-3/2}), \tag 3.17 $$ and $$ |g_{\pm}(\lambda, x)|+|g'_{\pm}(\lambda, x)|+|g''_{\pm}(\lambda, x)| \leq C, \quad x\lesseqgtr 0, \lambda\geq V_{+}+1, \tag 3.18 $$ where $$ g_{\pm}(\lambda, x)=e^{\mp ik_{\pm}(\lambda)x} f_{\pm}(\lambda, x). \tag 3.19 $$ (The estimate (3.18) immediately follows from (3.2).) Employing (3.15)--(3.19), one arrives at $$ W(\overline{f_{-}(\lambda)}, f_{+}(\lambda))\operatornamewithlimits{=} _{\lambda\to\infty} o(\lambda^{-1}) \tag 3.20 $$ after two integrations by parts in (3.14) (and a few tears) using the Riemann-Lebesgue lemma and $$\align g_{\pm}(\lambda, 0) &\operatornamewithlimits{=}_{\lambda\to\infty} 1\mp (1/2i\lambda^{1/2})\int\limits^{\pm\infty}_{0} dxe^{\mp ik_{\pm}x} [V(x)-V_{\pm}] f_{\pm}(\lambda, x)+O(\lambda^{-1}), \tag 3.21 \\ g'_{\pm}(\lambda, 0) &\operatornamewithlimits{=}_{\lambda\to\infty} O(\lambda^{-1/2}). \tag 3.22 \endalign $$ Combining (3.6)--(3.8), (3.12), and (3.20) then proves (3.11) \qed \enddemo The asymptotic behavior (3.11) slightly improves Lemma 1.4 (iv) in [4] since we arrive at the conclusion $o(\lambda^{-3/2})$ instead of $O(\lambda^{-3/2})$ and their extra hypothesis $\int\limits^{0}_{- \infty} dx(1+|x|)|V(x)|+\int\limits^{\infty}_{0}dx(1+|x|)|V(x)- V_{+}|<\infty$ is not needed in our proof. Remark 2.2 and the paragraph following it clearly apply in the present context. Lemma 3.2 enables one to again remove the Abelian limit in the trace formula (1.6) for $V(x)$. (We recall our notational conventions in (2.4)--(2.6), (2.8), (2.9), and the paragraph following (2.9).) \proclaim{Theorem 3.3} Suppose $V, V'\in AC_{\text{\rom{loc}}}(\Bbb R)$, $V$ is real-valued, $V', V''\in L^{1}(\Bbb R)$, $\int\limits^{0}_{- \infty} dx |V(x)|\mathbreak+\int\limits^{\infty}_{0}dx|V(x)-V_{+}|<\infty$ for some $V_{+}>0$. Let $E_{o}=\inf\,\sigma(H)$. Then $[1-2\xi(., x)]\in L^{1}((E_{o}, \infty); d\lambda)$, $x\in\Bbb R$ and $$\align V(x)&=E_{o}+\int\limits^{\infty}_{E_o}d\lambda[1-2\xi(\lambda, x)] \tag 3.23 \\ &=2\{e_{0}+\sum_{j\in J_{+}}[e_{j}-\mu_{j}(x)]\} + \int\limits^{\infty}_{0}d\lambda[1-2\xi(\lambda, x)] \tag 3.24 \\ &=2\{e_{0}+\sum_{j\in J_{+}}[e_{j}-\mu_{j}(x)]\} -(2/\pi)\int\limits^{\infty}_{0}d\lambda\text{\rom{ Im}}\left\{\ln \left[1+R^{\ell}(\lambda)\frac{f_{-}(\lambda, x)^{2}}{|f_{-}(\lambda, x)|^{2}}\right]\right\}, \, x\in\Bbb R. \tag 3.25 \endalign $$ If $\sigma_{p}(H)=\emptyset$, the discrete spectrum part $2\{\dots\}$ in \rom{(3.24)} and \rom{(3.25)} is to be deleted. \endproclaim Given Lemma 3.2, the proof of Theorem 3.3 is identical to that of Theorem 2.3. \remark{Remark \rom{3.4}} While $\xi(\lambda, x)$ is continuous in $\lambda >V_{+}$, $\xi(\lambda, x)$ may have (countably infinitely-many) jump discontinuities of size one in $[0, V_{+}]$. These discontinuities occur at those special energies $\mu_{j}(x)\in [0, V_{+}]$ which are eigenvalues of the Dirichlet operator $H^{D}_{+, x}$ (the restriction of $H^{D}_{x}$ to $(x, \infty)$). The Green's function $G(z, x, x)$ of $H$ is of the type $$ G(z, x, x)=[m^{-}_{x}(z)-m^{+}_{x}(z)]^{-1}, \tag 3.26 $$ where $m^{\pm}_{x}(z)$ are the Weyl $m$-functions associated with $H^{D}_{\pm, x}$ in $L^{2}((x, \pm\infty))(H^{D}_{x}=H^{D}_{-, x} \oplus H^{D}_{+, x})$ with $m^{-}_{x}(z)$ being continuous near $\mu_{j}(x)$ while $m^{+}_{x}(z)$ has a first-order pole with a negative residue at $\mu_{j}(x)$. Thus $$ G(z, x, x)\operatornamewithlimits{=}_{z\to\mu_{j}(x)} c[z-\mu_{j}(x)] +O([z-\mu(x)]^{2}), \quad \mu_{j}(x)\in (0, V_{+}) \tag 3.27 $$ for some $c>0$ and hence $\xi(\lambda, x)$ has the jump discontinuity $$ \lim_{\epsilon\downarrow 0} \xi(\mu_{j}(x)\pm\epsilon)=\cases 0 \\ 1 \endcases \tag 3.28 $$ at $\mu_{j}(x)\in (0, V_{+})$. A comparison of formulas (3.10) for $\xi(\lambda, x), \lambda >V_{-}$ and (3.7) for $R_{-}(\lambda)$ then shows that $$ 1+R^{\ell}(\mu_{j}(x))\frac{f_{-}(\mu_{j}(x), x)^{2}} {|f_{-}(\mu_{j}(x), x)|^{2}}=0 \tag 3.29 $$ since $f_{+}(\mu_{j}(x), x)=0$. Clearly $\xi(\lambda, x)$ is continuous in $\lambda >0$ away from these special energies $\mu_{j}(x)\in (0, V_{+}]$. \endremark Next, we shall briefly consider the behavior of $\xi(\lambda, x)$ as $\lambda\downarrow 0$ and as $\lambda\downarrow V_{+}$ (since $H$ changes spectral multiplicity at $V_{+}$) similarly to the discussion at the end of \S 2. We shall assume $$ \int\limits^{0}_{-\infty}dx (1+x^{2})|V(x)|+\int\limits^{\infty} _{0} dx (1+x^{2})|V(x)-V_{+}|<\infty \tag 3.30 $$ in addition to (3.1) and consider the following case distinctions depending on whether or not $H$ has a threshold resonance at $\lambda =0$. \example{Case I} $W(f_{-}(0), f_{+}(0))\neq 0$ and $f_{-}(0, x)f_{+}(0, x)\neq 0$. \endexample \flushpar Then, since $f_{\pm}(0, x)$ are real-valued, one infers $$ R^{\ell}(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow 0} -1+O(\lambda^{1/2}), \tag 3.31 $$ $$ G(\lambda +i0, x, x)\operatornamewithlimits{=}_{\lambda\downarrow 0} \frac{f_{+}(0, x)f_{-}(0, x)}{W(f_{+}(0), f_{-}(0))} + O(\lambda^{1/2}) \tag 3.32 $$ and hence $$ \xi(\lambda, x)=\pi^{-1}\arg[G(\lambda +i0, x, x)] \operatornamewithlimits{=}_{\lambda\downarrow 0} O(\lambda^{1/2}) \ \text{in case I}. \tag 3.33 $$ \example{Case II} $W(f_{-}(0), f_{+}(0))=0$ and $f_{-}(0, x)f_{+}(0, x)\neq 0$. \endexample \flushpar In this case one infers (see, e.g., Proposition 2.4 in [4] or Lemma 2.5 in [13]) that $$ W(f_{-}(\lambda), f_{+}(\lambda))\operatornamewithlimits{=}_ {\lambda\downarrow 0} i\gamma\lambda^{1/2}+O(\lambda), \quad \gamma\in\Bbb R\backslash\{0\}, \tag 3.34 $$ $$ G(z+i0, x, x)\operatornamewithlimits{=}_{\lambda\downarrow 0} (i/\gamma)f_{+}(0, x)f_{-}(0, x)[1+O(\lambda^{1/2})]. \tag 3.35 $$ Thus we get $$ \xi(\lambda, x)=\pi^{-1}\arg[G(\lambda +i0, x, x)] \operatornamewithlimits{=}_{\lambda\downarrow 0}\frac12 + O(\lambda^{1/2}) \ \text{in case II}. \tag 3.36 $$ Discussion of the case $\lambda\downarrow V_{+}$ remains. Since $f_{+}(V_{+}, x)$ is real-valued and $W(f_{-}(\lambda),\mathbreak f_{+}(\lambda))\neq 0$ for $\lambda > 0$ (see, e.g., Lemma 1.2 in [4] or Lemma 2.1 in [13]), one infers that $$ R^{r}(\lambda)\operatornamewithlimits{=}_{\lambda\downarrow V_{+}} -1+O((\lambda -V_{+})^{1/2}), \tag 3.37 $$ $$ G(\lambda +i0, x, x)\operatornamewithlimits{=}_{\lambda\downarrow V_{+}} \frac{f_{+}(V_{+}, x)f_{-}(V_{+}, x)}{W(f_{+}(V_{+}), f_{-}(V_{+}))} + O((\lambda - V_{+})^{1/2}), \tag 3.38 $$ and hence $$ \xi(\lambda, x)\operatornamewithlimits{=}_{\lambda\downarrow V_{+}} \frac12 + \pi^{-1}\text{Im}\left\{\ln\left[1+R^{\ell}(V_{+}) \frac{f_{-}(V_{+}, x)^{2}}{|f_{-}(V_{+}, x)|^{2}}\right]\right\} + O((\lambda -V_{+})^{1/2}). \tag 3.39 $$ This serves as an illustration that $\xi(\lambda, x)$ is insensitive to the fact that $H$ changes its spectral multiplicity at $V_{+}$. We conclude this section with a brief discussion of the case where $V(x)\operatornamewithlimits{\longrightarrow}\limits_{x\to\infty}\infty$, that is, we now assume $$ V\in C(\Bbb R), \quad \int\limits^{0}_{-\infty}dx|V(x)|<\infty, \quad \lim_{x\to\infty}V(x)=\infty. \tag 3.40 $$ $H$ is then defined as the form sum of $H_{o}=-\frac{d^2}{dx^2}$ and $V$ in $L^{2}(\Bbb R)$. Then $f_{-}(z, x)$ can be defined as in (3.2) (or (2.13)) and, since $V(x)\to\infty$ as $x\to\infty$, the Weyl $m$-function associated with $H^{D}_{+, 0}$ (the restriction of $H$ to $(0, \infty)$ with a Dirichlet boundary condition at $x=0$) is meromorphic. Hence, there exists an entire function $f_{+}(z, x)$ satisfying $$ f_{+}(z, .)\in L^{2}((0, \infty)), \quad z\in\Bbb C \tag 3.41 $$ and (3.3). $f_{+}(z, x)$ can be chosen to be real-valued for $\lambda\in\Bbb R$ (see, e.g., [26] for further details). $H$ now has simple spectrum which is purely absolutely continuous on $(0, \infty)$. The reflection coefficient from left incidence is then defined as in (3.4), that is, $$ R^{\ell}(\lambda)=-\frac{\overline{W(f_{-}(\lambda), f_{+}(\lambda))}} {W(f_{-}(\lambda), f_{+}(\lambda))}, \quad \lambda >0 \tag 3.42 $$ and hence $$ |R^{\ell}(\lambda)|=1, \quad \lambda > 0 \tag 3.43 $$ proves total reflection from left incidence at all positive energies $\lambda > 0$. The Green's function $G(z, x, x)$ of $H$ now satisfies $$\align G(\lambda +i0, x, x)&=\frac{f_{+}(\lambda, x)f_{-}(\lambda, x)} {W(f_{+}(\lambda), f_{-}(\lambda))}\\ &=(i/2\lambda^{1/2})|f_{-}(\lambda, x)|^{2}\left[1+R^{\ell}(\lambda) \frac{f_{-}(\lambda, x)^{2}}{|f_{-}(\lambda, x)|^{2}}\right], \quad \lambda > 0 \tag 3.44 \endalign $$ and one infers $$ \xi(\lambda, x)=\frac12 +\pi^{-1}\text{Im}\left\{\ln\left[1+R^{\ell} (\lambda)\frac{f_{-}(\lambda, x)^{2}}{|f_{-}(\lambda, x)|^{2}}\right] \right\}, \quad \lambda >0 \tag 3.45 $$ as in (3.10). However, due to the total reflection at all positive energies $\lambda >0$, $$ R^{\ell}(\lambda)=e^{ir(\lambda)}\operatornamewithlimits{\not\longrightarrow} _{\lambda\to\infty} 0 \tag 3.46 $$ for some real-valued function $r$ and hence $$ 1-2\xi(\lambda, x)\operatornamewithlimits{\not\longrightarrow}_ {\lambda\to\infty} 0. \tag 3.47 $$ In fact, in the explicit example $V(x)=e^{x}$ discussed, for instance in [26], where $R^{\ell}(\lambda)=- \exp\{2i\arg[\Gamma(1+2i\lambda^{1/2})]\}, \lambda\geq 0$ (here $\Gamma(.)$ denotes the gamma function), one can verify that $[1- 2\xi(., x)]\notin L^{1}((0, \infty); d\lambda), x\in\Bbb R$. As a consequence, the Abelian limit in the trace formula (1.6) for $V(x)$, in general, cannot be removed in the case (3.36) and hence (1.6) represents a genuine summability method in this situation. As mentioned briefly in the introduction, this becomes even more transparent in the case where $V(x)\operatornamewithlimits{\longrightarrow}\limits_{x\to\pm\infty}\infty$ since then for all $x\in\Bbb R$ and a.e.~$\lambda\in\Bbb R$, $|1-2\xi (\lambda, x)|=1$. \bigpagebreak \flushpar {\bf \S 4. Hill Operators with Impurities} In our final section, we shall consider short-range perturbations $W$ of Hill operators $H^{o}=\frac{d^2}{dx^2}+V^{o}$ and hence extend the results of \S2 to scattering off impurities (defects) in one-dimensional solids. Again, most of the results in this section are valid under minimal smoothness assumptions on $V^o$ and $W$. However, since our main result in Theorem 4.3 requires a certain regularity of $V^o$ and $W$, we shall avoid technicalities and suppose these regularity assumptions throughout this section. We start by briefly reviewing the necessary Floquet theory associated with the periodic background potential $V^o$ satisfying $$ V^{o}\in H^{1, 2}([0, a]), \quad V^{o} \ \text{real-valued}, \quad V^{o}(x+a)=V^{o}(x), \quad x\in\Bbb R \tag 4.1 $$ for some $a>0$. The corresponding Hill operator $H^o$ in $L^{2}(\Bbb R)$ is then defined by $$ H^{o}=-\frac{d^2}{dx^2}+V^{o}, \quad \Cal D(H^{o})=H^{2, 2}(\Bbb R). \tag 4.2 $$ The spectrum of $H^o$ is purely absolutely continuous of the type $$ \sigma(H^{o})=\bigcup_{n\in\Bbb N} [E^{o}_{2(n-1)}, E^{o}_{2n-1}], E^{o}_{0}0$ for $\lambda < E^{o}_0$ and hence $$\gathered -i\theta(\lambda)>0, \lambda 0$, $W\in H^{2,1}(\Bbb R)$ is real-valued, and $W\in L^{1}(\Bbb R; (1+|x|)dx)$. Then for all $x\in\Bbb R$, $$ \xi(\lambda, x)=0, \quad \lambda0$ for $z<\inf\,\sigma(H)$ and $G(z, x, x)$ is real-valued for $z$ in any (non-empty) spectral gap of $H$. Equation (4.44) follows from (2.24) and (4.39) (we note that $s^{o}(\lambda, a)$ is real-valued for $\lambda\in\Bbb R$ and $\theta(\lambda)\in\Bbb R$ for $\lambda\in\sigma(H^{o})$ by (4.10)). Since $G(\lambda +i0, x, x)$ is continuous and zero-free for $\lambda\in\sigma(H^{o})^{o}$, $\xi(\lambda, x)$ is continuous in $\lambda\in\sigma(H^{o})$. Inequality (4.45) is then clear from (4.44). For the explicit construction of the compact intervals $\sigma_n$ with $E^{o}_{n}\in\partial\sigma_{n}, \sum\limits_{n\in\Bbb N_{0}}|\sigma_{n}|<\infty$ such that $$ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} o(\lambda^{-1/2}) \ \text{for} \ \lambda\in\sigma(H^{o})\backslash \bigcup_{n\in\Bbb N_{0}}\sigma_{n}, \tag 4.50 $$ we refer to [9], [10]. Here we only mention that (4.50) is implied by the asymptotic relation (4.41) and, assuming $\lambda\in \sigma(H^{o})\backslash \operatornamewithlimits{\cup}\limits_{n\in\Bbb N_{0}}\sigma_{n}$, by $$ W(f^{o}_{-}(\lambda), f^{o}_{+}(\lambda))= \frac{2i\sin[\theta(\lambda)a]}{s^{o}(\lambda, a)} \operatornamewithlimits{=}_{\lambda\to\infty} 2i\lambda^{1/2}[1+O(|\lambda |^{-1/8})], \tag 4.51 $$ $$ T(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} 1+(1/2i\lambda^{1/2})\left[\int\limits_{\Bbb R}dxW(x)+ O(|\lambda |^{-1/8})\right], \tag 4.52 $$ $$ R\Sp r\\ \ell\endSp(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} (1/2i\lambda^{1/2})\left[\int\limits_{\Bbb R} dxW(x)e^{\mp 2i\theta(\lambda)x} +O(|\lambda |^{-1/8})\right] \tag 4.53 $$ as proven in [9] (see also [10]). In particular, in order to arrive at (4.52), (4.53) one combines (4.41), (4.51), (4.36)--(4.38), and $$ p_{\pm}(\lambda, x)\operatornamewithlimits{=}_{\lambda\to\infty} 1+O(|\lambda |^{-1/8}), \quad \lambda\in\sigma(H^{o})\backslash \bigcup_{n\in\Bbb N_{0}}\sigma_{n}, \tag 4.54 $$ $$ |f_{\pm}(\lambda, x)|\leq C(1+|x|), \quad \lambda\in\sigma(H^{o}), \tag 4.55 $$ $$ |f_{\pm}(\lambda, x)-f^{o}_{\pm}(\lambda, x)|\leq C(1+|x|)(1+|\lambda |)^{-1/2}, \quad \lambda\in\sigma(H^{o}). \tag 4.56 $$ Here $p_{\pm}(\lambda, x)$ has been introduced in (4.11), and (4.55) and (4.56) follow from (4.32), (4.33) (see [9], [10]). Analogous relations for the first and second $x$-derivatives of $p_{\pm}(\lambda, x)$ and $f_{\pm}(\lambda, x)$ then yield $$ R^{r(\ell)}(\lambda)\operatornamewithlimits{=}_{\lambda\to\infty} o(\lambda^{-3/2}), \quad \lambda\in\sigma(H^{o})\backslash \bigcup_{n\in\Bbb N_{0}}\sigma_{n} \tag 4.57 $$ after two integrations by parts in (4.37), (4.38). Together with (4.45) this proves (4.46). Using $|R^{r(\ell)}(\lambda)|\leq 1, \lambda\in\sigma(H^{o})$ (as a consequence of the unitarity of the scattering matrix (4.35)), one then infers (4.47) from $$ \int\limits_{\sigma(H^{o})} d\lambda |R^{r(\ell)}(\lambda)| \leq \int\limits_{\operatornamewithlimits{\cup}\limits_{n\in\Bbb N_{0}} \sigma_{n}}d\lambda + \int\limits_{\sigma(H^{o})\backslash \operatornamewithlimits{\cup}\limits_{n\in\Bbb N_{0}}\sigma_{n}} |o(\lambda^{-3/2})|<\infty. \qed \tag 4.58 $$ \enddemo As in the previous sections, Lemma 4.2 will enable us to remove the Abelian limit in the trace formula (2.27) for $V(x)$ and state the principal result of this section. (We recall our notational conventions in (4.3), (4.17), (4.25)--(4.31).) \proclaim{Theorem 4.3} Suppose $V=V^{o}+W$ where $V^{o}\in H^{1, 2}([0, a])$ is real-valued, $V^{o}(x+a)=V^{o}(x)$ for some $a>0$, $W\in H^{2, 1}(\Bbb R)$ is real-valued, and $W\in L^{1}(\Bbb R; (1+|x|)dx)$. Let $E_{o}=\inf\,\sigma(H)$. Then $[1-2\xi(., x)]\in L^{1}((E_{o}, \infty); d\lambda), x\in\Bbb R$ and $$\align V(x)&=V^{o}(x)+W(x)=E_{o}+\int\limits^{\infty}_{E_o}d\lambda [1-2\xi(\lambda, x)] \tag 4.59 \\ &=\{2e_{0, 0}+2\sum_{j\in J_{0, +}}[e_{0, j}-\mu_{0, j}(x)]- E^{o}_{0}\} \\ &+ \sum^{\infty}_{n=1} \{E^{o}_{2n-1}+ 2\sum_{j\in J_{n, +}} [e_{n, j}-\mu_{n, j}(x)]-E^{o}_{2n}\} +\int\limits_{\sigma(H^{o})} d\lambda[1-2\xi(\lambda, x)] \tag 4.60 \\ &=\{2e_{0, 0}+2\sum_{j\in J_{0, +}} [e_{0, j}-\mu_{0, j}(x)]- E^{o}_{0}\} \\ &+ \sum^{\infty}_{n=1}\{E^{o}_{2n-1}+2\sum_{j\in J_{n, +}} [e_{n, j}-\mu_{n, j}(x)]-E^{o}_{2n}\} \\ &- (2/\pi)\int\limits_{\sigma(H^{o})} d\lambda\text{\rom{ Im}}\left\{\ln\left[ 1+R\Sp r\\ \ell\endSp(\lambda)\frac{f_{\pm}(\lambda, x)^{2}} {|f_{\pm}(\lambda, x)|^{2}}\right]\right\}, \quad x\in\Bbb R. \tag 4.61 \endalign $$ Similarly, $$\align W(x)&=2\{e_{0, 0}+\sum_{j\in J_{0, +}}[e_{0, j}-\mu_{0, j}(x)]- E^{o}_{0}\} \\ &+2\sum^{\infty}_{n=1}\{\mu^{o}_{n}(x) +\sum_{j\in J_{0, +}}[e_{n, j}-\mu_{n, j}(x)]-E^{o}_{2n}\} +\int\limits_{\sigma(H^{o})} d\lambda [1-2\xi(\lambda, x)] \tag 4.62 \\ &=2\{e_{0, 0}+\sum_{j\in J_{0, +}}[e_{0, j}-\mu_{0, j}(x)]- E^{o}_{0}\} \\ &+2\sum^{\infty}_{n=1}\{\mu^{o}_{n}(x)+\sum_{j\in J_{n, +}} [e_{n, j}-\mu_{n, j}(x)]-E^{o}_{2n}\} \\ &-(2/\pi)\int\limits_{\sigma(H^{o})} d\lambda\text{\rom{ Im}}\left\{\ln \left[1+R\Sp r\\ \ell\endSp(\lambda)\frac{f_{\pm}(\lambda, x)^{2}}{|f_{\pm} (\lambda, x)|^{2}}\right]\right\}, \quad x\in\Bbb R. \tag 4.63 \endalign $$ If $\sigma_{p, n}(H)=\emptyset$, the corresponding expression $\{\dots\}$ in \rom{(4.60)}, \rom{(4.61)} is to be replaced by $\{E^{o}_{2n-1}-2\mu_{n, 0}(x)+E^{o}_{2n}\}$ if $n\in\Bbb N$ and deleted if $n=0$. Similarly, if $\sigma_{p, n}(H)=\emptyset$, the corresponding expression $2\{\dots\}$in \rom{(4.62)}, \rom{(4.63)} is to be replaced by $2\{\mu^{o}_{n}(x)-\mu_{n, 0}(x)\}$ if $n\in\Bbb N$ and deleted if $n=0$. \endproclaim \demo{Proof} The trace formula (4.53) follows from (1.6), (4.45) (4.47), and the Lebesgue dominated convergence theorem. Equalities (4.54) and (4.55) are then clear from (4.59), (4.44), (4.48), and (4.49). Equations (4.62) and (4.63) are obtained by combining (4.60), (4.61) and (4.19) observing the finite total gap length (4.21). \qed \enddemo We note that the analog of Remark 2.2 clearly holds in the present context. Moreover, the threshold behavior of $\xi(\lambda, x)$ in (2.43)--(2.51) near $\lambda=0$ extends to the essential spectral band edges $\{E^{o}_{n}\}_{n\in\Bbb N_{0}}$ of $H$ in the current impurity scattering situation. In particular, assuming $$ W\in L^{1}(\Bbb R; (1+x^{2})dx) \tag 4.64 $$ in addition to (4.22), one again distinguishes two cases depending on whether or not $H$ has a threshold resonance at $E^{o}_{n}$. \example{Case I} $W(f_{-}(E^{o}_{n}), f_{+}(E^{o}_{n}))\neq 0$ and $f_{-}(E^{o}_{n}, x)f_{+}(E^{o}_{n}, x)\neq 0$. \endexample \flushpar Then $$ R^{r(\ell)}(E^{o}_{n})=-1 \tag 4.65 $$ and $$ \xi(\lambda, x)\operatornamewithlimits{=}\Sb \lambda\to E^{o}_{n}\\ \lambda\in\sigma (H^{o}) \endSb O(|\lambda -E^{o}_{n}|^{1/2}) \ \text{in case I}. \tag 4.66 $$ \example{Case II} $W(f_{-}(E^{o}_{n}), f_{+}(E^{o}_{n}))=0$ and $f_{-}(E^{o}_{n},x)f_{+}(E^{o}_{n}, x)\neq 0$. \endexample \flushpar Then one can show that $$ \xi(\lambda, x)\operatornamewithlimits{=}\Sb \lambda\to E^{o}_{n}\\ \lambda\in\sigma (H^{o}) \endSb \frac12 + O(|\lambda - E^{o}_{n}|^{1/2}) \ \text{in case II}. \tag 4.67 $$ \vskip 0.4in \example{Acknowledgments} We would like to thank Z.~Zhao for numerous helpful discussions. F.G.~is indebted to the Department of Mathematical Sciences of the University of Trondheim, Norway, and the Department of Mathematics at Caltech, Pasadena, CA for the hospitality extended to him in the summer of 1993. Support by the Norwegian Research Council for Science and Humanities (NAVF) and by Caltech is gratefully acknowledged. H.H.~is grateful to the Norwegian Research Council for Science and the Humanities (NAVF) for support. \endexample \vskip 0.4in \Refs \endRefs \item{[1]} N.~Aronszajn and W.F.~Donoghue, {\it On exponential representations of analytic functions in the upper half-plane with positive imaginary part}, J.~Anal.~Math. {\bf 5} (1957), 321--388. \item{[2]} D.~Boll\'e, F.~Gesztesy, and S.F.J.~Wilk, {\it A complete treatment of low-energy scattering in one dimension}, J.~Operator Theory {\bf 13} (1985), 3--31. \item{[3]} D.~Boll\'e, F.~Gesztesy, and M.~Klaus, {\it Scattering theory for one-dimensional systems with $\int dxV(x)=0$}, J.~Math.~Anal.~Appl. {\bf 122} (1987), 496--518. \item{[4]} A.~Cohen and T.~Kappeler, {\it Scattering and inverse scattering for steplike potentials in the Schr\"odinger equation}, Indiana Univ.~Math.~J. {\bf 34} (1985), 127--180. \item{[5]} W.~Craig, {\it The trace formula for Schr\"odinger operators on the line}, Commun.~Math. Phys. {\bf 126} (1989), 379--407. \item{[6]} E.B.~Davies and B.~Simon, {\it Scattering theory for systems with different spatial asymptotics to the left and right}, Commun.~Math.~Phys. {\bf 63} (1978), 277--301. \item{[7]} P.~Deift and E.~Trubowitz, {\it Inverse scattering on the line}, Commun.~Pure Appl.~Math. {\bf 32} (1979), 121--251. \item{[8]} B.A.~Dubrovin, {\it Periodic problems for the Korteweg-de Vries equation in the class of finite band potentials}, Funct.~Anal.~Appl. {\bf 9} (1975), 215--223. \item{[9]} N.E.~Firsova, {\it Levinson formula for perturbed Hill operator}, Theoret.~Math.~Phys. {\bf 62} (1985), 130--140. \item{[10]} N.E.~Firsova, {\it The direct and inverse scattering problems for the one-dimensional perturbed Hill operator}, Math.~USSR Sbornik {\bf 58} (1987), 351--388. \item{[11]} H.~Flaschka, {\it On the inverse problem for Hill's operator}, Arch.~Rat.~Mech.~Anal. {\bf 59} (1975), 293--309. \item{[12]} I.M.~Gel'fand and B.M.~Levitan, {\it On a simple identity for eigenvalues of a second order differential operator}, Dokl.~Akad.~Nauk SSSR {\bf 88} (1953), 593--596 (Russian); English transl.~in Izrail M.~Gelfand, Collected Papers Vol. I (S.G.~Gindikin, V.W.~Guillemin, A.A.~Kirillov, B.~Kostant, S.~Sternberg, eds.) Springer, Berlin, 1987, pp. 457--461. \item{[13]} F.~Gesztesy, {\it Scattering theory for one-dimensional systems with nontrivial spatial\linebreak asymptotics}, Schr\"odinger Operators, Aarhus 1985 (E. Balslev, ed.), Lecture Notes in Math. Vol. 1218, Springer, Berlin, 1986, pp. 93--122. \item{[14]} F.~Gesztesy, {\it New trace formulas for Schr\"odinger operators}, Evolution Equations (G.~Ferreyra, G.~Goldstein, F.~Neubrander, eds.), Marcel Dekker, to appear. \item{[15]} F.~Gesztesy and B.~Simon, {\it A short proof of Zheludev's theorem}, Trans.~Amer.~Math.~Soc. {\bf 335} (1993), 329--340. \item{[16]} F.~Gesztesy and B.~Simon, {\it The xi function}, to be submitted to Ann. Math. \item{[17]} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it A trace formula for multidimensional Schr\"odinger operators}, to be submitted to Invent.~Math. \item{[18]} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it Trace formulae and inverse spectral theory for Schr\"odinger operators,} Bull.~Amer.~Math.~Soc. {\bf 29} (1993), 250--255. \item{[19]} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it Higher order trace relations for Schr\"o-dinger operators}, to be submitted to Commun.~Pure Appl.~Math. \item{[20]} F.~Gesztesy, W. Karwowski, and Z.~Zhao, {\it New types of soliton solutions}, Bull.~Amer. Math.~Soc. {\bf 27} (1992), 266--272. \item{[21]} F.~Gesztesy, W. Karwowski, and Z.~Zhao, {\it Limits of soliton solutions}, Duke Math.~J. {\bf 68} (1992), 101--150. \item{[22]} H. Hochstadt, {\it On the determination of a Hill's equation from its spectrum}, Arch.~Rat. Mech.~Anal. {\bf 19} (1965), 353--362. \item{[23]} K.~Iwasaki, {\it Inverse problem for Sturm-Liouville and Hill equation}, Ann. Mat. Pura Appl. Ser. 4, {\bf 149} (1987), 185--206. \item{[24]} S.~Kotani and M.~Krishna, {\it Almost periodicity of some random potentials}, J.~Funct.~Anal. {\bf 78} (1988), 390--405. \item{[25]} M.G.~Krein, {\it Perturbation determinants and a formula for the traces of unitary and self-adjoint operators}, Sov.~Math.~Dokl. {\bf 3} (1962), 707--710. \item{[26]} G.~Kristensson, {\it The one-dimensional inverse scattering problem for an increasing potential}, J.~Math. Phys. {\bf 27} (1986), 804--815. \item{[27]} B.M.~Levitan, {\it On the closure of the set of finite-zone potentials}, Math.~USSR Sbornik {\bf 51} (1985), 67--89. \item{[28]} B.M.~Levitan, {\it Inverse Sturm-Liouville Problems}, VNU Science Press, Utrecht, 1987. \item{[29]} V.A.~Marchenko, {\it Sturm-Liouville Operators and Applications}, Birkh\"auser, Basel, 1986. \item{[30]} H.P.~McKean and P.~van Moerbeke, {\it The spectrum of Hill's equation}, Invent.~Math. {\bf 30} (1975), 217--274. \item{[31]} F.S.~Rofe-Beketov, {\it A test for the finiteness of the number of discrete levels introduced into gaps of a continuous spectrum by perturbations of a periodic potential}, Sov.~Math. Dokl. {\bf 5} (1964), 689--692. \item{[32]} F.S.~Rofe-Beketov, {\it Perturbation of a Hill operator having a first moment and nonzero integral creates one discrete level in distant spectral gaps}, Mat.~Fizika i Funkts.~Analiz (Khar'kov) {\bf 19} (1973), 158--159 (Russian). \item{[33]} B.~Simon, {\it Spectral analysis of rank one perturbations and applications}, Lecture given at the 1993 Vancouver Summer School, preprint. \item{[34]} E.~Trubowitz, {\it The inverse problem for periodic potentials}, Commun.~Pure Appl.~Math. {\bf 30} (1977), 321--337. \item{[35]} S.~Venakides, {\it The infinite period limit of the inverse formalism for periodic potentials}, Commun.~Pure Appl.~Math. {\bf 41} (1988), 3--17. \item{[36]} V.A.~Zheludev, {\it Eigenvalues of the perturbed Schroedinger operator with a periodic potential}, Topics in Mathematical Physics (M.Sh.~Birman, ed.), Vol. 2, Consultants Bureau, New York, 1968, 87--101. \item{[37]} V.A.~Zheludev, {\it Perturbation of the spectrum of the one-dimensional self-adjoint Schr\"o-dinger operator with a periodic potential}, Topics in Mathematical Physics (M.Sh.~Birman, ed.), Vol. 4, Consultants Bureau, New York, 1971, 55--75. \enddocument