Content-Type: multipart/mixed; boundary="-------------0207241512621" This is a multi-part message in MIME format. ---------------0207241512621 Content-Type: text/plain; name="02-321.keywords" Content-Transfer-Encoding: 7bit Content-Disposition: attachment; filename="02-321.keywords" Ergodicity, Bernoulli property, hyperbolicity, mathematical billiards, cylindric billiards, singularity ---------------0207241512621 Content-Type: application/x-tex; name="dis.tex" Content-Transfer-Encoding: 7bit Content-Disposition: inline; filename="dis.tex" \input amstex.tex \documentstyle{amsppt} \magnification=\magstep1 \vsize 8.5truein \hsize 6truein %\NoBlackBoxes %\TagsOnRight % {\catcode`\@=11\gdef\logo@{}} \define\flow{\left(\bold{M},\{S^t\}_{t\in\Bbb R},\mu\right)} \define\symb{\Sigma=\left(\sigma_1,\sigma_2,\dots,\sigma_n\right)} \define\traj{S^{[a,b]}x_0} \define\endrem{} \noindent July 24, 2002 \bigskip \bigskip \heading Proving The Ergodic Hypothesis for Billiards \\ With Disjoint Cylindric Scatterers \endheading \bigskip \bigskip \centerline{{\bf N\'andor Sim\'anyi} \footnote{Research supported by the National Science Foundation, grant DMS-0098773 and by the Hungarian National Foundation for Scientific Research, grants OTKA-26176 and OTKA-29849.}} \bigskip \bigskip \centerline{University of Alabama at Birmingham} \centerline{Department of Mathematics} \centerline{Campbell Hall, Birmingham, AL 35294 U.S.A.} \centerline{E-mail: simanyi\@math.uab.edu} \bigskip \bigskip \hbox{\centerline{\vbox{\hsize 8cm {\bf Abstract.} In this paper we study the ergodic properties of mathematical billiards describing the uniform motion of a point in a flat torus from which finitely many, pairwise disjoint, tubular neighborhoods of translated subtori (the so called cylindric scatterers) have been removed. We prove that every such system is ergodic (actually, a Bernoulli flow), unless a simple geometric obstacle for the ergodicity is present.}}} \bigskip \bigskip \noindent Primary subject classification: 37D50 \medskip \noindent Secondary subject classification: 34D05 \bigskip \bigskip \bigskip \heading \S1. Introduction \endheading \bigskip \bigskip Non-uniformly hyperbolic systems (possibly, with singularities) play a pivotal role in the ergodic theory of dynamical systems. Their systematic study started several decades ago, and it is not our goal here to provide the reader with a comprehensive review of the history of these investigations but, instead, we opt for presenting in nutshell a cross section of a few selected results. In 1939 G. A. Hedlund and E. Hopf [He(1939)], [Ho(1939)], proved the hyperbolic ergodicity of geodesic flows on closed, compact surfaces with constant negative curvature by inventing the famous method of "Hopf chains" constituted by local stable and unstable invariant manifolds. In 1963 Ya. G. Sinai [Sin(1963)] formulated a modern version of Boltzmann's ergodic hypothesis, what we call now the "Boltzmann-Sinai ergodic hypothesis": the billiard system of $N$ ($\ge2$) hard spheres of unit mass moving in the flat torus $\Bbb T^\nu=\Bbb R^\nu/\Bbb Z^\nu$ ($\nu\ge2$) is ergodic after we make the standard reductions by fixing the values of the trivial invariant quantities. It took seven years until he proved this conjecture for the case $N=2$, $\nu=2$ in [Sin(1970)]. Another 17 years later N. I. Chernov and Ya. G. Sinai [S-Ch(1987)] proved the hypothesis for the case $N=2$, $\nu\ge2$ by also proving a powerful and very useful theorem on local ergodicity. In the meantime, in 1977, Ya. Pesin [P(1977)] laid down the foundations of his theory on the ergodic properties of smooth, hyperbolic dynamical systems. Later on this theory (nowadays called Pesin theory) was significantly extended by A. Katok and J-M. Strelcyn [K-S(1986)] to hyperbolic systems with singularities. That theory is already applicable for billiard systems, too. Until the end of the seventies the phenomenon of hyperbolicity (exponential instability of the trajectories) was almost exclusively attributed to some direct geometric scattering effect, like negative curvature of space, or strict convexity of the scatterers. This explains the profound shock that was caused by the discovery of L. A. Bunimovich [B(1979)]: certain focusing billiard tables (like the celebrated stadium) can also produce complete hyperbolicity and, in that way, ergodicity. It was partly this result that led to Wojtkowski's theory of invariant cone fields, [W(1985)], [W(1986)]. The big difference between the system of two spheres in $\Bbb T^\nu$ ($\nu\ge2$, [S-Ch(1987)]) and the system of $N$ ($\ge3$) spheres in $\Bbb T^\nu$ is that the latter one is merely a so called semi-dispersive billiard system (the scatterers are convex but not strictly convex sets, namely cylinders), while the former one is strictly dispersive (the scatterers are strictly convex sets). This fact makes the proof of ergodicity (mixing properties) much more complicated. In our series of papers jointly written with A. Kr\'amli and D. Sz\'asz [K-S-Sz(1990)], [K-S-Sz(1991)], and [K-S-Sz(1992)], we managed to prove the (hyperbolic) ergodicity of three and four billiard spheres in the toroidal container $\Bbb T^\nu$. By inventing new topological methods and the Connecting Path Formula (CPF), in my two-part paper [Sim(1992)] I proved the (hyperbolic) ergodicity of $N$ hard spheres in $\Bbb T^\nu$, provided that $N\le\nu$. The common feature of hard sphere systems is --- as D. Sz\'asz pointed this out first in [Sz(1993)] and [Sz(1994)] --- that all of them belong to the family of so called cylindric billiards, the definition of which can be found later in this paragraph. However, the first appearance of a special, 3-D cylindric billiard system took place in [K-S-Sz(1989)], where we proved the ergodicity of a 3-D billiard flow with two orthogonal cylindric scatterers. Later D. Sz\'asz [Sz(1994)] presented a complete picture (as far as ergodicity is concerned) of cylindric billiards with cylinders whose generator subspaces are spanned by mutually orthogonal coordinate axes. The task of proving ergodicity for the first non-trivial, non-orthogonal cylindric billiard system was taken up in [S-Sz(1994)]. Finally, in our joint venture with D. Sz\'asz [S-Sz(1999)] we managed to prove the complete hyperbolicity of {\it typical} hard sphere systems. \subheading{\bf Cylindric billiards} Consider the $d$-dimensional ($d\ge2$) flat torus $\Bbb T^d=\Bbb R^d/\Cal L$ supplied with the usual Riemannian inner product $\langle\, .\, ,\, .\, \rangle$ inherited from the standard inner product of the universal covering space $\Bbb R^d$. Here $\Cal L\subset\Bbb R^d$ is supposed to be a lattice, i. e. a discrete subgroup of the additive group $\Bbb R^d$ with $\text{rank}(\Cal L)=d$. The reason why we want to allow general lattices, other than just the integer lattice $\Bbb Z^d$, is that otherwise the hard sphere systems would not be covered. The geometry of the structure lattice $\Cal L$ in the case of a hard sphere system is significantly different from the geometry of the standard lattice $\Bbb Z^d$ in the standard Euclidean space $\Bbb R^d$, see subsection 2.4 of [Sim(2002)]. The configuration space of a cylindric billiard is $\bold Q=\Bbb T^d\setminus\left(C_1\cup\dots\cup C_k\right)$, where the cylindric scatterers $C_i$ ($i=1,\dots,k$) are defined as follows: Let $A_i\subset\Bbb R^d$ be a so called lattice subspace of $\Bbb R^d$, which means that $\text{rank}(A_i\cap\Cal L)=\text{dim}A_i$. In this case the factor $A_i/(A_i\cap\Cal L)$ is a subtorus in $\Bbb T^d=\Bbb R^d/\Cal L$ which will be taken as the generator of the cylinder $C_i\subset\Bbb T^d$, $i=1,\dots,k$. Denote by $L_i=A_i^\perp$ the orthocomplement of $A_i$ in $\Bbb R^d$. Throughout this paper we will always assume that $\text{dim}L_i\ge2$. Let, furthermore, the numbers $r_i>0$ (the radii of the spherical cylinders $C_i$) and some translation vectors $t_i\in\Bbb T^d=\Bbb R^d/\Cal L$ be given. The translation vectors $t_i$ play a crucial role in positioning the cylinders $C_i$ in the ambient torus $\Bbb T^d$. Set $$ C_i=\left\{x\in\Bbb T^d:\; \text{dist}\left(x-t_i,A_i/(A_i\cap\Cal L) \right)0$ is the common radius of the disks, while $m_1,\dots,m_N$ are the masses.) \medskip \subheading{\bf Theorem of [Sim(2002)]} Every hard sphere system is completely hyperbolic, i. e. all of its relevant Lyapunov exponents are nonzero almost everywhere. \medskip In this paper we are mainly interested in understanding the ergodic properties of cylindric billiard flows $\flow$ in which the closures $\bar C_i$ of the scattering cylinders $C_i$ are pairwise disjoint, i. e. $$ \bar C_i\cap\bar C_j=\emptyset\text{ for }1\le i0) \; \text{s. t.} \; \forall \alpha \in (-\delta,\delta) \\ V\left(S^a\left(Q(x)+\alpha W,V(x)\right)\right)=V(S^ax)\text{ and } V\left(S^b\left(Q(x)+\alpha W,V(x)\right)\right)=V(S^bx)\big\}. \endaligned $$ \endproclaim \endrem ($\Cal Z$ is the common tangent space $\Cal T_q\bold Q$ of the parallelizable manifold $\bold Q$ at any of its points $q$, while $V(x)$ is the velocity component of the phase point $x=\left(Q(x),\,V(x)\right)$.) It is known (see (3) in \S 3 of [S-Ch (1987)]) that $\Cal N_0(S^{[a,b]}x)$ is a linear subspace of $\Cal Z$ indeed, and $V(x)\in \Cal N_0(S^{[a,b]}x)$. The neutral space $\Cal N_t(S^{[a,b]}x)$ of the segment $S^{[a,b]}x$ at time $t\in [a,b]$ is defined as follows: $$ \Cal N_t(S^{[a,b]}x)=\Cal N_0\left(S^{[a-t,b-t]}(S^tx)\right). $$ It is clear that the neutral space $\Cal N_t(S^{[a,b]}x)$ can be canonically identified with $\Cal N_0(S^{[a,b]}x)$ by the usual identification of the tangent spaces of $\bold Q$ along the trajectory $S^{(-\infty,\infty)}x$ (see, for instance, \S 2 of [K-S-Sz(1990)]). Our next definition is that of the {\bf advance}. Consider a non-singular orbit segment $S^{[a,b]}x$ with the symbolic collision sequence $\Sigma=(\sigma_1, \dots, \sigma_n)$ ($n\ge 1$), meaning that $S^{[a,b]}x$ has exactly $n$ collisions with $\partial\bold Q$, and the $i$-th collision ($1\le i\le n$) takes place at the boundary of the cylinder $C_{\sigma_i}$. For $x=(Q,V)\in\bold M$ and $W\in\Cal Z$, $\Vert W\Vert$ sufficiently small, denote $T_W(Q,V):=(Q+W,V)$. \proclaim{Definition 2.2} For any $1\le k\le n$ and $t\in[a,b]$, the advance $$ \alpha(\sigma_k)\colon\;\Cal N_t(S^{[a,b]}x) \rightarrow \Bbb R $$ of the collision $\sigma_k$ is the unique linear extension of the linear functional $\alpha(\sigma_k)$ defined in a sufficiently small neighborhood of the origin of $\Cal N_t(S^{[a,b]}x)$ in the following way: $$ \alpha(\sigma_k)(W):= t_k(x)-t_k(S^{-t}T_WS^tx). $$ \endproclaim \endrem Here $t_k=t_k(x)$ is the time moment of the $k$-th collision $\sigma_k$ on the trajectory of $x$ after time $t=a$. The above formula and the notion of the advance functional $$ \alpha_k=\alpha(\sigma_k):\; \Cal N_t\left(S^{[a,b]}x\right)\to\Bbb R $$ has two important features: \medskip (i) If the spatial translation $(Q,V)\mapsto(Q+W,V)$ is carried out at time $t$, then $t_k$ changes linearly in $W$, and it takes place just $\alpha_k(W)$ units of time earlier. (This is why it is called ``advance''.) \medskip (ii) If the considered reference time $t$ is somewhere between $t_{k-1}$ and $t_k$, then the neutrality of $W$ precisely means that $$ W-\alpha_k(W)\cdot V(x)\in A_{\sigma_k}, $$ i. e. a neutral (with respect to the collision $\sigma_k$) spatial translation $W$ with the advance $\alpha_k(W)=0$ means that the vector $W$ belongs to the generator space $A_{\sigma_k}$ of the cylinder $C_{\sigma_k}$. It is now time to bring up the basic notion of {\bf sufficiency} (or, sometimes it is also called {\bf geometric hyperbolicity}) of a trajectory (segment). This is the utmost important necessary condition for the proof of the fundamental theorem for semi-dispersive billiards, see Condition (ii) of Theorem 3.6 and Definition 2.12 in [K-S-Sz(1990)]. \medskip \proclaim{Definition 2.3} \roster \item The nonsingular trajectory segment $S^{[a,b]}x$ ($a$ and $b$ are supposed not to be moments of collision) is said to be {\bf sufficient} if and only if the dimension of $\Cal N_t(S^{[a,b]}x)$ ($t\in [a,b]$) is minimal, i.e. $\text{dim}\ \Cal N_t(S^{[a,b]}x)=1$. \item The trajectory segment $S^{[a,b]}x$ containing exactly one singularity (a so called ``simple singularity'', see above) is said to be {\bf sufficient} if and only if both branches of this trajectory segment are sufficient. \endroster \endproclaim \endrem \medskip \proclaim{Definition 2.4} The phase point $x\in\bold M$ with at most one singularity is said to be sufficient if and only if its whole trajectory $S^{(-\infty,\infty)}x$ is sufficient, which means, by definition, that some of its bounded segments $S^{[a,b]}x$ are sufficient. \endproclaim \endrem In the case of an orbit $S^{(-\infty,\infty)}x$ with a simple singularity, sufficiency means that both branches of $S^{(-\infty,\infty)}x$ are sufficient. \bigskip \subheading{No accumulation (of collisions) in finite time} By the results of Vaserstein [V(1979)], Galperin [G(1981)] and Burago-Ferleger-Kononenko [B-F-K(1998)], in a semi-dis\-per\-sive billiard flow there can only be finitely many collisions in finite time intervals, see Theorem 1 in [B-F-K(1998)]. Thus, the dynamics is well defined as long as the trajectory does not hit more than one boundary components at the same time. \bigskip \subheading{Slim sets} We are going to summarize the basic properties of codimension-two subsets $A$ of a smooth manifold $M$. Since these subsets $A$ are just those negligible in our dynamical discussions, we shall call them {\bf slim}. As to a broader exposition of the issues, see [E(1978)] or \S 2 of [K-S-Sz(1991)]. Note that the dimension $\dim A$ of a separable metric space $A$ is one of the three classical notions of topological dimension: the covering (\v Cech-Lebesgue), the small inductive (Menger-Urysohn), or the large inductive (Brouwer-\v Cech) dimension. As it is known from general general topology, all of them are the same for separable metric spaces. \medskip \proclaim{Definition 2.5} A subset $A$ of $M$ is called slim if and only if $A$ can be covered by a countable family of codimension-two (i. e. at least two) closed sets of $\mu$--measure zero, where $\mu$ is a smooth measure on $M$. (Cf. Definition 2.12 of [K-S-Sz(1991)].) \endproclaim \endrem \medskip \proclaim{Property 2.6} The collection of all slim subsets of $M$ is a $\sigma$-ideal, that is, countable unions of slim sets and arbitrary subsets of slim sets are also slim. \endproclaim \medskip \proclaim{Proposition 2.7 (Locality)} A subset $A\subset M$ is slim if and only if for every $x\in A$ there exists an open neighborhood $U$ of $x$ in $M$ such that $U\cap A$ is slim. (Cf. Lemma 2.14 of [K-S-Sz(1991)].) \endproclaim \medskip \proclaim{Property 2.8} A closed subset $A\subset M$ is slim if and only if $\mu(A)=0$ and $\dim A\le\dim M-2$. \endproclaim \medskip \proclaim{Property 2.9 (Integrability)} If $A\subset M_1\times M_2$ is a closed subset of the product of two manifolds, and for every $x\in M_1$ the set $$ A_x=\{ y\in M_2\colon\; (x,y)\in A\} $$ is slim in $M_2$, then $A$ is slim in $M_1\times M_2$. \endproclaim \medskip The following propositions characterize the codimension-one and codimension-two sets. \proclaim{Proposition 2.10} For any closed subset $S\subset M$ the following three conditions are equivalent: \roster \item"{(i)}" $\dim S\le\dim M-2$; \item"{(ii)}" $\text{int}S=\emptyset$ and for every open connected set $G\subset M$ the difference set $G\setminus S$ is also connected; \item"{(iii)}" $\text{int}S=\emptyset$ and for every point $x\in M$ and for any open neighborhood $V$ of $x$ in $M$ there exists a smaller open neighborhood $W\subset V$ of the point $x$ such that for every pair of points $y,z\in W\setminus S$ there is a continuous curve $\gamma$ in the set $V\setminus S$ connecting the points $y$ and $z$. \endroster \endproclaim \noindent (See Theorem 1.8.13 and Problem 1.8.E of [E(1978)].) \medskip \proclaim{Proposition 2.11} For any subset $S\subset M$ the condition $\dim S\le\dim M-1$ is equivalent to $\text{int}S=\emptyset$. (See Theorem 1.8.10 of [E(1978)].) \endproclaim \medskip We recall an elementary, but important lemma (Lemma 4.15 of [K-S-Sz(1991)]). Let $R_2$ be the set of phase points $x\in\bold M\setminus\partial\bold M$ such that the trajectory $S^{(-\infty,\infty)}x$ has more than one singularities. \proclaim{Proposition 2.12} The set $R_2$ is a countable union of codimension-two smooth sub-manifolds of $M$ and, being such, it is slim. \endproclaim \medskip The next lemma establishes the most important property of slim sets which gives us the fundamental geometric tool to connect the open ergodic components of billiard flows. \proclaim{Proposition 2.13} If $M$ is connected, then the complement $M\setminus A$ of a slim set $A\subset M$ necessarily contains an arc-wise connected, $G_\delta$ set of full measure. (See Property 3 of \S 4.1 in [K-S-Sz(1989)]. The $G_\delta$ sets are, by definition, the countable intersections of open sets.) \endproclaim \medskip \subheading{\bf The subsets $\bold M^0$ and $\bold M^\#$} Denote by $\bold M^\#$ the set of all phase points $x\in\bold M$ for which the trajectory of $x$ encounters infinitely many non-tangential collisions in both time directions. The trajectories of the points $x\in\bold M\setminus\bold M^\#$ are lines: the motion is linear and uniform, see the appendix of [Sz(1994)]. It is proven in lemmas A.2.1 and A.2.2 of [Sz(1994)] that the closed set $\bold M\setminus\bold M^\#$ is a finite union of hyperplanes. It is also proven in [Sz(1994)] that, locally, the two sides of a hyperplanar component of $\bold M\setminus\bold M^\#$ can be connected by a positively measured beam of trajectories, hence, from the point of view of ergodicity, in this paper it is enough to show that the connected components of $\bold M^\#$ entirely belong to one ergodic component. This is what we are going to do in this paper. Denote by $\bold M^0$ the set of all phase points $x\in\bold M^\#$ the trajectory of which does not hit any singularity, and use the notation $\bold M^1$ for the set of all phase points $x\in\bold M^\#$ whose orbit contains exactly one, simple singularity. According to Proposition 2.12, the set $\bold M^\#\setminus(\bold M^0\cup\bold M^1)$ is a countable union of smooth, codimension-two ($\ge2$) submanifolds of $\bold M$, and, therefore, this set may be discarded in our study of ergodicity, please see also the properties of slim sets above. Thus, we will restrict our attention to the phase points $x\in\bold M^0\cup\bold M^1$. \medskip \subheading{\bf The ``Chernov-Sinai Ansatz''} An essential precondition for the Theorem on Local Ergodicity by B\'alint--Chernov--Sz\'asz--T\'oth is the so called ``Chernov-Sinai Ansatz'' which we are going to formulate below. Denote by $\Cal S\Cal R^+\subset\partial\bold M$ the set of all phase points $x_0=(q_0,v_0)\in\partial\bold M$ corresponding to singular reflections (a tangential or a double collision at time zero) supplied with the post-collision (outgoing) velocity $v_0$. It is well known that $\Cal S\Cal R^+$ is a compact cell complex with dimension $2d-3=\text{dim}\bold M-2$. It is also known (see Lemma 4.1 in [K-S-Sz(1990)]) that for $\nu$-almost every phase point $x_0\in\Cal S\Cal R^+$ (Here $\nu$ is the Riemannian volume of $\Cal S\Cal R^+$ induced by the restriction of the natural Riemannian metric of $\bold M$.) the forward orbit $S^{(0,\infty)}x_0$ does not hit any further singularity. The Chernov-Sinai Ansatz postulates that for $\nu$-almost every $x_0\in\Cal S\Cal R^+$ the forward orbit $S^{(0,\infty)}x_0$ is sufficient (geometrically hyperbolic). \medskip \subheading{\bf The Theorem on Local Ergodicity} The Theorem on Local Ergodicity by B\'alint--Chernov--Sz\'asz--T\'oth (Theorem 4.4 of [B-Ch-Sz-T(2001)]) claims the following: Let $\flow$ be a semi-dispersive billiard flow with the properties (1.1)--(1.2) and such that the smooth components of the boundary $\partial\bold Q$ of the configuration space are algebraic hypersurfaces. (The cylindric billiards with (1.1)--(1.2) automatically fulfill this algebraicity condition.) Assume -- further -- that the Chernov-Sinai Ansatz holds true, and a phase point $x_0\in\bold M\setminus\partial\bold M$ is given with the properties \medskip (i) $S^{(-\infty,\infty)}x$ has at most one singularity, \noindent and (ii) $S^{(-\infty,\infty)}x$ is sufficient. (In the case of a singular obit $S^{(-\infty,\infty)}x$ this means that both branches of $S^{(-\infty,\infty)}x$ are sufficient.) \medskip Then some open neighborhood $U_0\subset\bold M$ of $x_0$ belongs to a single ergodic component of the flow $\flow$. (Modulo the zero sets, of course.) \bigskip \bigskip \heading \S3. Geometric Considerations \endheading \bigskip \bigskip Consider a non-singular trajectory segment $S^{[a,b]}x_0=\left\{x_t=S^tx_0\big|\; a\le t\le b\right\}$ of the cylindric billiard flow $\flow$ with the symbolic collision sequence $\symb$, meaning that there are time moments $a0$) of this normal vector as time $t$ elapses. If there is no collision on the orbit segment $S^{[0,t]}y$, then the relationship between $(\delta q,\, \delta v)\in\Cal T_y\bold M$ and $(\delta q',\, \delta v')=\left(DS^t\right)(\delta q,\, \delta v)$ is obviously $$ \aligned \delta v'&=\delta v, \\ \delta q'&=\delta q+t\delta v, \endaligned \tag 3.5 $$ from which we obtain that $$ \aligned (\delta q',\, \delta v')\in\Cal T_{y'}J&\Leftrightarrow\langle\delta q'-t \delta v',\, z\rangle+\langle\delta v',\, w\rangle=0 \\ &\Leftrightarrow \langle\delta q',\, z\rangle+\langle\delta v',\, w-tz\rangle=0. \endaligned $$ This means that $n_t=(z,\, w-tz)$. It is always very useful to consider the quadratic form $Q(n)=Q((z,w))=:\langle z,w\rangle$ associated with the normal vector $n=(z,w)\in\Cal T_y\bold M$ of $J$ at $y$. $Q(n)$ is the so called ``infinitesimal Lyapunov function'', see [K-B(1994)] or part A.4 of the Appendix in [Ch(1994)]. For a detailed exposition of the relationship between the quadratic form $Q$, the relevant symplectic geometry and the dynamics, please see [L-W(1995)]. \medskip \subheading{\bf Remark} Since the normal vector $n=(z,w)$ of $J$ is only determined up to a nonzero scalar multiplier, the value $Q(n)$ is only determined up to a positive multiplier. However, this means that the sign of $Q(n)$ (which is the utmost important thing for us) is uniquely determined. This remark will gain a particular importance in the near future. \medskip \noindent From the above calculations we get that $$ Q(n_t)=Q(n_0)-t||z||^2\le Q(n_0). \tag 3.6 $$ The next question is how the normal vector $n$ of $J$ gets transformed $n^-\mapsto n^+$ through a collision (reflection) at time $t=0$? Elementary geometric considerations show (see Lemma 2 of [Sin(1979)], or formula (2) in \S3 of [S-Ch(1987)]) that the linearization of the flow $$ \left(DS^t\right)\Big|_{t=0}:\; (\delta q^-,\, \delta v^-)\longmapsto (\delta q^+,\, \delta v^+) $$ is given by the formulae $$ \aligned \delta q^+&=R\delta q^-, \\ \delta v^+&=R\delta v^-+2\cos\phi RV^*KV\delta q^-, \endaligned \tag 3.7 $$ where the operator $R:\; \Cal T_q\bold Q\to \Cal T_q\bold Q$ is the orthogonal reflection across the tangent hyperplane $\Cal T_q\partial\bold Q$ of $\partial\bold Q$ at $q\in \partial\bold Q$ ($y^-=(q,v^-)\in\partial\bold M$, $y^+=(q,v^+)\in\partial\bold M$), $V:\; (v^-)^\perp\to\Cal T_q\partial\bold Q$ is the $v^-$-parallel projection of the orthocomplement hyperplane $(v^-)^\perp$ onto $\Cal T_q\partial\bold Q$, $V^*:\; \Cal T_q\partial\bold Q\to (v^-)^\perp$ is the adjoint of $V$, i. e. it is the projection of $\Cal T_q\partial\bold Q$ onto $(v^-)^\perp$ being parallel to the normal vector $\nu(q)$ of $\partial\bold Q$ at $q\in\partial\bold Q$, $K:\; \Cal T_q\partial\bold Q\to \Cal T_q\partial\bold Q$ is the second fundamental form of $\partial\bold Q$ at $q$ and, finally, $\cos\phi=\langle\nu(q),\, v^+\rangle$ is the cosine of the angle $\phi$ subtended by $v^+$ and the normal vector $\nu(q)$. For the formula (3.7), please also see the last displayed formula of \S1 in [S-Ch(1982)], or (i) and (ii) of Proposition 2.3 in [K-S-Sz(1990)]. We note that it is enough to deal with the tangent vectors $(\delta q^-,\, \delta v^-)\in(v^-)^\perp\times(v^-)^\perp$ ($(\delta q^+,\, \delta v^+)\in(v^+)^\perp\times(v^+)^\perp$), for the manifold $J$ under investigation is supposed to be flow-invariant, so any vector $(\delta q,\, \delta v)=(\alpha v,\, 0)$ ($\alpha\in\Bbb R$) is automatically inside $\Cal T_yJ$. The backward version (inverse) $$ \left(DS^t\right)\Big|_{t=0}:\; (\delta q^+,\, \delta v^+)\mapsto (\delta q^-,\, \delta v^-) $$ can be deduced easily from (3.7): $$ \aligned \delta q^-&=R\delta q^+, \\ \delta v^-&=R\delta v^+-2\cos\phi RV_1^*KV_1\delta q^+, \endaligned \tag 3.8 $$ where $V_1:\; (v^+)^\perp\to\Cal T_q\partial\bold Q$ is the $v^+$-parallel projection of $(v^+)^\perp$ onto $\Cal T_q\partial\bold Q$. By using formula (3.8), one easily computes the time-evolution $n^-\longmapsto n^+$ of a normal vector $n^-=(z,w)\in\Cal T_{y^-}\bold M$ of $J$ if a collision $y^-\longmapsto y^+$ takes place at time $t=0$: $$ \aligned (\delta q^+,\, \delta v^+)\in\Cal T_{y^+}J\Leftrightarrow\langle R\delta q^+, \, z\rangle+\langle R\delta v^+-2\cos\phi RV_1^*KV_1\delta q^+,\, w\rangle &=0 \\ \Leftrightarrow\langle\delta q^+,\, Rz-2\cos\phi V_1^*KV_1Rw\rangle+ \langle\delta v^+,\, Rw\rangle &=0. \endaligned $$ This means that $$ n^+=\left(Rz-2\cos\phi V_1^*KV_1Rw,\, Rw\right) \tag 3.9 $$ if $n^-=(z,\, w)$. It follows that $$ \aligned Q(n^+)&=Q(n^-)-2\cos\phi\langle V_1^*KV_1Rw,\, Rw\rangle \\ &=Q(n^-)-2\cos\phi\langle KV_1Rw,\, V_1Rw\rangle\le Q(n^-). \endaligned \tag 3.10 $$ Here we used the fact that the second fundamental form $K$ of $\partial\bold Q$ at $q$ is positive semi-definite, which just means that the billiard system is semi-dispersive. The last simple observation on the quadratic form $Q(n)$ regards the involution $I:\; \bold M\to\bold M$, $I(q,v)=(q,-v)$ corresponding to the time reversal. If $n=(z,w)$ is a normal vector of $J$ at $y$, then, obviously, $I(n)=(z,-w)$ is a normal vector of $I(J)$ at $I(y)$ and $$ Q\left(I(n)\right)=-Q(n). \tag 3.11 $$ By switching --- if necessary --- from the separating manifold $J$ to $I(J)$, and by taking a suitable remote image $S^t(J)$ ($t>>1$), in the spirit of (3.6), (3.10)--(3.11) we can assume that $$ Q(n)<0 \tag 3.12 $$ for every {\it unit} normal vector $n\in\Cal T_y\bold M$ of $J$ near a phase point $y\in J$. \medskip \subheading{\bf Remark 3.13} There could be, however, a little difficulty in achieving the inequality $Q(n)<0$, i. e. (3.12). Namely, it may happen that $Q(n_t)=0$ for every $t\in\Bbb R$. According to (3.6), the equation $Q(n_t)=0$ ($\forall\, t\in\Bbb R$) implies that $n_t=:(z_t,\, w_t)=(0,\, w_t)$ for all $t\in\Bbb R$ and, moreover, in the view of (3.9), $w_t^+=Rw_t^-$ is the transformation law at any collision $y_t=(q_t,\, v_t)\in\partial\bold M$. Furthermore, at every collision $y_t=(q_t,\, v_t)\in\partial\bold M$ the projected tangent vector $V_1Rw_t^-=V_1w_t^+$ lies in the null space of the operator $K$ (see also (3.9)), and this means that $w_0$ is a neutral vector for the entire trajectory $S^{\Bbb R}y$, i. e. $w_0\in\Cal N\left(S^{\Bbb R}y\right)$. (For the notion of neutral vectors and $\Cal N\left(S^{\Bbb R}y\right)$, cf. \S2 above.) On the other hand, this is impossible for the following reason: Any tangent vector $(\delta q,\delta v)$ from the space $\Cal N\left(S^{\Bbb R}y\right)\times\Cal N\left(S^{\Bbb R}y\right)$ is automatically tangent to the separating manifold $J$, thus for any normal vector $n=(z,w)\in\Cal T_y\bold M$ of a separating manifold $J$ one has $$ (z,\, w)\in\Cal N\left(S^{\Bbb R}y\right)^\perp\times\Cal N\left( S^{\Bbb R}y\right)^\perp. \tag 3.14 $$ (As a direct inspection shows. We always tacitly assume that the exceptional manifold $J$ is locally defined by the equation $J=\left\{x\in U_0\big|\;\text{ dim}\Cal N\left(S^{[a,b]}x\right)>1\right\}$ with orbit segments $S^{[a,b]}x$ whose symbolic sequence is combinatorially rich, i. e. it typically provides sufficient phase points.) The membership in (3.14) is, however, impossible with a nonzero vector $w\in\Cal N\left(S^{\Bbb R}y\right)$. \qed \medskip \subheading{\bf Singularities} \medskip Consider a smooth, connected piece $\Cal S\subset\bold M$ of a singularity manifold corresponding to a singular (tangential or double) reflection {\it in the past}. Such a manifold $\Cal S$ is locally flow-invariant and has one codimension, so we can speak about its normal vectors $n$ and the uniquely determined sign of $Q(n)$ for $0\ne n\in\Cal T_y\bold M$, $y\in\Cal S$, $n\perp\Cal S$ (depending on the foot point, of course). Consider first a phase point $y^+\in\partial\bold M$ right after the singular reflection that is described by $\Cal S$. It follows from the proof of Lemma 4.1 of [K-S-Sz(1990)] and Sub-lemma 4.4 therein that at $y^+=(q,\, v^+)\in\partial\bold M$ any tangent vector $(0,\, \delta v)\in\Cal T_{y^+}\bold M$ lies actually in $\Cal T_{y^+}\Cal S$ and, consequently, the normal vector $n=(z,w)\in\Cal T_{y^+}\bold M$ of $\Cal S$ at $y^+$ necessarily has the form $n=(z,0)$, i. e. $w=0$. Thus $Q(n)=0$ for any normal vector $n\in\Cal T_{y^+}\bold M$ of $\Cal S$. According to the monotonicity inequalities (3.6) and (3.10) above, $$ Q(n)<0 \tag 3.15 $$ for any phase point $y\in\Cal S$ of a past singularity manifold $\Cal S$. \medskip The above observations lead to the following conclusion: \medskip \subheading{\bf Proposition 3.16} Assume that the separating manifold $J\subset\bold M$ ($J$ is smooth, connected, $\text{codim}(J)=1$) is selected in such a way that $Q(n_y)<0$ for all normal vectors $0\ne n_y\in\Cal T_y\bold M$ of $J$ at any point $y\in J$, see above. Suppose further that the non-singular orbit segments $S^{[a,b]}y$ ($y\in B_0$, $B_0$ is a small open ball, $0b$. \medskip \subheading{\bf Proof} As we have seen before, the manifold $E$ is defined by the relation $v_t\in\text{span}\left\{A_{\sigma_i},\, A_{\sigma_{i+1}}\right\}$ with some $i\in\{1,2,\dots,n-1\}$, $\text{dim}\left(\text{span}\left\{A_{\sigma_i},\, A_{\sigma_{i+1}}\right\} \right)=d-1$, $t(\sigma_i)b$. Then, by the non-increasing property of $Q(n_\tau)$ in $\tau$, there is a small $\epsilon>0$ such that $Q(n_\tau)=0$ for all $\tau$, $t(\sigma_{i+1})<\tau0$, i. e. $E^{(1)}$ and $E^{(2)}$ are transversal at any point of their intersection. This finishes the proof of the corollary. \qed \bigskip \bigskip \heading \S4. Hyperbolicity Is Abundant \\ The Inductive Proof \endheading \bigskip \bigskip Below we present the inductive proof of the Theorem of this paper. The induction will be performed with respect to the number of cylinders $k$. Beside the ergodicity (and, therefore, the Bernoulli property, see [C-H(1996)]) and [O-W(1998)]) we will prove (and use as the induction hypothesis!) a few technical properties listed below: \medskip (H1) The Chernov--Sinai Ansatz (see \S2 above) holds true for the cylindric billiard flow $\flow$; \medskip (H2) There exists a slim subset $S\subset\bold M$ (see \S2 for the concept of ``slimness'') such that for all $x\in\bold M\setminus S$ \medskip \hskip 0.3truein (i) $S^{(-\infty,\infty)}x$ has at most one singularity and \hskip 0.3truein (ii) $S^{(-\infty,\infty)}x$ is hyperbolic (in the singular case both branches of $S^{(-\infty,\infty)}x$ are supposed to be hyperbolic, see \S2 above). \medskip Consequently, according to the Fundamental Theorem for algebraic, semi-dis\-persive billiards (Theorem 4.4 of [B-Ch-Sz-T(2001)]), \medskip (H3) For every $x\in\bold M\setminus S$ the assertion of Theorem 4.4 of [B-Ch-Sz-T(2001)] holds true in some open neighborhood $U_0$ of $x$ in $\bold M$, in particular, $x$ is a so called ``zig-zag point'', see Definition 5.1 in [Sz(2000)]. Consequently, since the complementer set $\bold M\setminus S$ is known to contain a connected set of full measure (see \S2) and the open neighborhood $U_0$ of $x$ belongs to a single ergodic component, we get that \medskip (H4) $\flow$ is ergodic, hence it is a Bernoulli flow by [C-H(1996)] and [O-W(1998)]. \medskip The above properties (H1)---(H2) will serve for us as the induction hypothesis. \bigskip \heading 1. The base of the induction: $k=1$ \endheading \bigskip In this case, necessarily, $L_1=\Bbb R^d$ and $A_1=\{0\}$, so the cylindric billiard system is actually a genuine, $d$-dimensional Sinai--billiard with a single spherical scatterer which has been well known to enjoy the properties (H1)---(H2) since the seminal work [S-Ch(1987)]. \bigskip \heading 2. The induction step: $0$. There are two important facts here: \medskip (A) The weakly stable manifolds $\gamma^{ws}(x)$ (yet to be constructed for typical $x\in A_1$) are concave, local orthogonal sub-manifolds (see the ``Invariant Manifolds'' part of \S2 in [K-S-Sz(1990)]) and, as such, they are uniformly transversal to the manifold $\Cal S_0$, see Sub-lemma 4.2 in [K-S-Sz(1990)]; \medskip (B) The exponentially stable part $$ \aligned \gamma^{es}(x)=\bigg\{y=(q_1+\delta q_1+q_2,\, v_1+\delta v_1+v_2)\bigg| \\ \text{dist}\left(S^t_*(q_1,v_1),\, S^t_*(q_1+\delta q_1,v_1+\delta v_1)\right)\to 0 \\ \text{exp. fast as }t\to\infty,\text{ and } ||\delta q_1||+||\delta v_1||<\epsilon_0\bigg\} \endaligned \tag 4.6/a $$ of $\gamma^{ws}(x)$ ($x=(q_1+q_2,v_1+v_2)\in A_1$) is to be constructed by using the statement of the Fundamental Theorem (Theorem 4.4 of [B-Ch-Sz-T(2001)]) for the $\Cal C_0$-sub-billiard system $\left\{S^t_*\right\}$. This statement can be used, for the $\nu$-typical phase points $x=(q,v)=(q_1+q_2,v_1+v_2)$ of $A_1$ have the property that the $S_*$-part $\left\{S^t_*(q_1,v_1)\right\}$ of their forward orbit is hyperbolic with respect to the sub-billiard system defined by the cylinders $C_i$, $i\in\Cal C_0$, see Corollary 3.18. \medskip According to the above points (A) and (B), there exists a measurable subset $A_2\subset A_1$ with $\nu(A_2)>0$ and a number $\delta_0>0$ such that for every $x\in A_2$ the manifold $\gamma^{ws}(x)$ exists and its boundary is at least at the distance $\delta_0$ from $x$ (these distances are now measured by using the induced Riemannian metric on $\gamma^{ws}(x)$). Then, by the absolute continuity of the foliation, see Theorem 4.1 in [K-S(1986)], the union $$ B_2=\bigcup_{x\in A_2}\gamma^{ws}(x)\subset\bold M $$ has a positive $\mu$-measure in the phase space $\bold M$. Finally, the genuine forward orbits $S^{(0,\infty)}x$ of all points $x\in A_2$ avoid a fixed open ball $B_{r_0}$ of radius $r_0>0$. (For example: We may take any open ball $B_{r_0}$ inside the interior of any avoided cylinder $C_j$, $j\not\in\Cal C_0$.) Therefore, the forward orbit in the direct product dynamics $\left(S^t_*\times T^t_*\right)(y)$ of any point $y\in B_2$ ($y\in\gamma^{ws}(x)$, $x\in A_2$) avoids a slightly shrunk open ball $B_{r_0-\delta_0}$ of reduced radius $r_0-\delta_0$. However, this is clearly impossible, for the following reason: For $y=(q_1+q_2,\, v_1+v_2)$ ($q_1\in\tilde L$, $q_2\in\tilde A$, $v_1\in L^*$, $v_2\in A^*$) the $v_2$ component is left invariant by the product flow $S^t_*\times T^t_*$, and for almost every fixed value $v_2\in A^*$ (namely, for those vectors $v_2$ for which the orbit $tv_2/\left(A^*\cap\Bbb Z^d\right)$ ($t\in\Bbb R$) is dense in the torus $\tilde A=A^*/\left(A^*\cap\Bbb Z^d\right)$) the product flow $S^t_*\times T^t_*$ is ergodic on the corresponding level set, since it is the product of a mixing and an ergodic flow. The obtained contradiction finishes the indirect proof of the Chernov-Sinai Ansatz, that is, (H1). \qed \bigskip \subheading{\bf Corollary 4.7} The set $$ NH(\Cal S_0)=\left\{x\in\Cal S_0\big|\; S^{(0,\infty)}x \text{ is not hyperbolic}\right\} $$ is a slim set. (In the case of a singular forward orbit non-hyperbolicity of $S^{(0,\infty)}x$ is meant that at least one branch of $S^{(0,\infty)}x$ is not hyperbolic, see \S2.) \medskip \subheading{\bf Proof} Since the complement set $\Cal S_0\setminus A=\Cal S_0\setminus A_0$ is a countable union of smooth, proper sub-manifolds of $\Cal S_0$, the set $\Cal S_0\setminus A_0$ is slim. Therefore, it is enough to prove that the intersection $NH(\Cal S_0)\cap A_0$ is slim. However, according to Corollary 3.18, the forward orbit $S^{(0,\infty)}x$ of every $x\in A_0$ is hyperbolic, unless $x$ belongs to a countable union of smooth, proper sub-manifolds of $\Cal S_0$. Thus $NH(\Cal S_0)\cap A_0$ is slim. \qed \medskip In view of Lemma 4.1 of [K-S-Sz(1990)], the set $R_2$ of phase points with more than one singularity on their orbit is slim, see also \S2. Therefore, the final step in proving the remaining unproven induction hypothesis (i. e. (H2)) for our considered model $\flow$ with $k$ ($\ge2$) cylindric scatterers is to show that the set $$ D=\left\{x\in\bold M^0\setminus\partial\bold M \big|\; S^{(-\infty,\infty)}x \text{ is not hyperbolic}\right\} \tag 4.8 $$ is slim, i. e. it can be covered by a countable collection of closed subsets $F\subset\bold M$ with $\mu(F)=0$ and $\text{dim}F\le\text{dim}\bold M-2$. By the locality of slimness, see \S2 above, it is enough to prove that for every element $x\in D$ the point $x$ has an open neighborhood $U$ (in $\bold M$) such that the set $U\cap D$ is slim. We want to classify the phase points $x\in D$. Consider, therefore, an arbitrary phase point $x=(q,v)\in D$. Denote the doubly infinite, symbolic collision sequence of $S^{(-\infty,\infty)}x$ by $\Sigma=\left(\dots,\sigma_{-2},\sigma_{-1},\,\sigma_{1},\sigma_{2},\dots\right)$ so that $\sigma_1$ is the first collision in positive time. (The index $0$ is not used.) We distinguish between two cases: \medskip \subheading{Case I} $L^*=\text{span}\left\{L_{\sigma_i}|\; i\in\Bbb Z\setminus\{0\}\right\}\ne\Bbb R^d$. In this case, as we have seen before, the dynamics of $S^{(-\infty,\infty)}x$ is finitely covered by the direct product flow $\left\{S^t_*\times T^t_*\right\}$, where $\left\{S^t_*\right\}$ is the cylindric billiard flow in the sub-torus $\tilde L=L^*/\left(L^*\cap\Bbb Z^d\right)$ with the scatterers $C_{\sigma_i}\cap\tilde L$, while $\left\{T^t_*\right\}$ is the almost periodic (uniform) motion in the orthocomplement torus $\tilde A=A^*/\left(A^*\cap\Bbb Z^d\right)$, $A^*=(L^*)^\perp$. Now the point is that for the cylindric billiard flow $(\tilde L,\left\{S^t_*\right\},\mu_{\tilde L})$ both of the induction hypotheses (H1)--(H2) and, consequently, Theorem 5.2 of [Sz(2000)] apply. For the phase point $x\in D$ the direct product flow $\left(S^t_*\times T^t_*\right)(x)$ avoids an open ball, namely any open ball in the interior of any avoided cylinder $C_j$ with $$ j\not\in\left\{\sigma_i|\; i\in\Bbb Z\setminus\{0\}\right\}. $$ Consequently, for each component $(q_2,v_2)\in\tilde A\times A^*$ of the canonical decomposition of $x=(q,v)=(q_1+q_2,v_1+v_2)$, $q_1\in\tilde L$, $v_1\in L^*$, $q_2\in\tilde A$, $v_2\in A^*$ it is true that the $\tilde L$-orbit $S^t_*(q_1,v_1)$ of $(q_1,v_1)$ avoids an open set $\emptyset\ne B\subset\tilde L$ on a doubly unbounded set $H$ of time moments, $\inf H=-\infty$, $\sup H=+\infty$. Therefore, in view of Theorem 5.2 of [Sz(2000)], the $(q_1,v_1)$-part of the phase point $x=(q_1+q_2,v_1+v_2)$ belongs to a slim subset $S_1$ of the phase space $\tilde L\times L^*$. According to the integrability property of closed slim sets (cf. Property 4 in \S4.1 of [K-S-Sz(1989)]), even the closure $\bar D_1$ of the set $$ D_1=\left\{x\in D\big|\; \text{span}\left\{L_{\sigma_i(x)}|\; i\in\Bbb Z \setminus\{0\}\right\}\ne\Bbb R^d\right\} \tag 4.9 $$ (covered by Case I) is a slim subset of the phase space $\bold M$. We note that the set $\bar D_1$ is contained in the closed zero-set $$ \aligned K=\big\{x\in\bold M^\#\big|\; x \text{ has a trajectory branch with a symbolic sequence } \\ (\dots,\, \sigma_{-1},\, \sigma_1,\, \dots) \text{ such that }\text{span}\left\{L_{\sigma_i}|\; i\in\Bbb Z\setminus\{0\} \right\}\ne\Bbb R^d\big\}, \endaligned $$ and the argument with ``integrating up'' the closed slim sets (by using Property 4 in \S4.1 of [K-S-Sz(1989)]) is applied to the closed set $K$. \bigskip \subheading{Case II} $L^*=\text{span}\left\{L_{\sigma_i}|\; i\in\Bbb Z\setminus\{0\}\right\}=\Bbb R^d$. Select a vector $0\ne w\in\Cal N\left(S^{(-\infty,\infty)}x\right)$, $w\perp v$, from the neutral space $$ \Cal N\left(S^{(-\infty,\infty)}x\right)=\Cal N(x) $$ of the considered phase point $x\in D$. For $i\in\Bbb Z\setminus\{0\}$ denote by $\alpha_i=\alpha_i(w)$ the ``advance'' of the collision $\sigma_i$ corresponding to the neutral vector $w$, see \S2. Since $w$ is not parallel to $v$, at least two advances with neighboring indices are unequal; we may assume that $\alpha_{-1}\ne\alpha_1$. It follows from the proof of Lemma 3.2 that the event $\alpha_{-1}\ne\alpha_1$ can only occur if $$ v=v_0\in\text{span}\left\{A_{\sigma_{-1}},\, A_{\sigma_{1}}\right\}. \tag 4.10 $$ If the event $\alpha_k\ne\alpha_{k+1}$ ($k\ne-1,\, 0$) took place for another pair of neighboring advances as well, then, again by the proof of Lemma 3.2, we would have $$ v_t\in\text{span}\left\{A_{\sigma_{k}},\, A_{\sigma_{k+1}}\right\} \quad (t_kt_1$, $Q(n_\tau)>0$ for $\taut_{k+1}$, and $Q(\tilde n_\tau)>0$ for $\tau0} A_{\sigma_l}=\left(\text{span}\left\{L_{\sigma_l}\big|\; l>0\right\}\right)^\perp, \tag 4.14 $$ see also the closing part of the proof of Proposition 3.1. An analogous argument shows that $$ w-\alpha_{-1}v \in\bigcap_{k<0} A_{\sigma_k}=\left(\text{span}\left\{L_{\sigma_k}\big|\; k<0\right\}\right)^\perp. \tag 4.15 $$ The equations (4.14)--(4.15) and $\alpha_{-1}\ne0$ imply that $$ \aligned v\in\text{span}\left\{\bigcap_{k<0}A_{\sigma_k},\, \bigcap_{l>0}A_{\sigma_l}\right\} \\ =\bigcap_{k<0}A_{\sigma_k}+\bigcap_{l>0}A_{\sigma_l}:=H. \endaligned \tag 4.16 $$ Recall that $\bigcap_{n\ne0}A_{\sigma_n}=\{0\}$ in the actual Case II, and $H\ne\Bbb R^d$, since $\text{span}\allowmathbreak\left\{A_{\sigma_k},\, A_{\sigma_l}\right\}\ne \Bbb R^d$ for $k<00\right\}=\Cal B\big\} \endaligned \tag 4.17 $$ (with given $\Cal A,\, \Cal B\subset\{1,2,\dots,k\}$ such that $\text{span}\left\{L_j|\; j\in\Cal A\cup\Cal B\right\}=\Bbb R^d$) of the considered type again decompose as $(q,v)=(q_1+q_2,\, v_1+v_2)$, $v_1\in L^*=\text{span}\allowmathbreak\left\{L_{\sigma_l}|\; l>0\right\}$, $v_2\in A^*=\left(L^*\right)^\perp=\bigcap_{l>0}A_{\sigma_l}$, $q_1\in\tilde L=L^*/\left(L^*\cap\Bbb Z^d\right)$, $q_2\in\tilde A=A^*/\left(A^*\cap\Bbb Z^d\right)$, and the forward orbit $S^{(0,\infty)}x$ of our considered phase point $x\in D\cap J$ (fulfilling all of the mentioned assumptions) is essentially (up-to a finite covering) is governed by the product flow $\left(q_1(t),v_1(t)\right)=S^t_*(q_1,v_1)$, $\left(q_2(t),v_2(t)\right)=T^t_*(q_2,v_2)=(q_2+tv_2,v_2)$, where (as said before) $S^t_*$ is the sub-billiard flow in $\tilde L$ defined by the intersections of the cylinders $\left\{C_{\sigma_l}|\; l>0\right\}$ with the torus $\tilde L$. \medskip \subheading{\bf Lemma 4.18} The exponentially stable component $\gamma^{es}(x)$ of $\gamma^{ws}(x)$ ($x\in D\cap J$) defined by (4.6/a) is transversal to the codimension-one manifold $J$ described by the membership in (4.16). \medskip \subheading{\bf Proof} Argue by contradiction. Assume that $\Cal T_x\gamma^{es}(x)\subset\Cal T_xJ$. The tangent space $\Cal T_xJ$ is obviously given by the simple formula $$ \Cal T_xJ=\left\{(\delta q,\delta v)\in\Cal T_x\bold M\big|\; \delta v\in H \right\}. \tag 4.19 $$ The second fundamental form $B\left(\gamma^{es}(x)\right)$ of $\gamma^{es}(x)$ at the phase point $x=(q_1+q_2,v_1+v_2)$ is known to be negative definite, so its range is the entire orthocomplement $(v_1)^\perp$ of $v_1$ in the space $$ L^*=\text{span}\left\{L_{\sigma_l}|\; l>0\right\}= \left(\bigcap_{l>0}A_{\sigma_l}\right)^\perp. $$ On the other hand, since $$ v=v_1+v_2\in H=\bigcap_{k<0}A_{\sigma_k}+\bigcap_{l>0}A_{\sigma_l} $$ and $v_2\in A^*=\bigcap_{l>0}A_{\sigma_l}\subset H$, from the assumed relation $\Cal T_x\gamma^{es}(x)\subset\Cal T_xJ$ and from $v_1\in H$ we get that $L^*\subset H$. Since $A^*=\left(L^*\right)^\perp\subset H$, this means that $H=\Bbb R^d$, contradicting $\text{dim}H=d-1$. This finishes the proof of the lemma. \qed \medskip Finally, the slimness of the set $D$ in (4.8) will be proven in Case II as soon as we show that $\nu_J(\bar D)=0$, where $\bar D=\bar D(\Cal A,\Cal B)$ is defined in (4.17). This is, however, obtained the same way as the relation $\nu(A_2)=0$ at the end of the proof of the Ansatz. Indeed, in the case $\nu_J(\bar D)>0$ the union $$ \tilde D:=\bigcup_{x\in\bar D}\gamma^{es}(x) $$ would have a positive $\mu$-measure in $\bold M$ (by the transversality proved above and by the absolute continuity of the $\gamma^{es}(\,.\,)$ foliation, see Theorem 4.1 in [K-S(1986)]), but this is impossible, for all forward orbits $S^{(0,\infty)}y$ of the points $y\in\tilde D$ would avoid a common open ball that can be obtained by slightly shrinking any open ball inside the interior of any avoided cylinder $C_j$ with $j\not\in\Cal B$, see also the closing part of the proof of $\nu(A_2)=0$ above. This finishes the proof of the fact that the set $D$ in (4.8) is indeed slim. From this, from the proved Chernov-Sinai Ansatz, and from the quoted slimness of the set $R_2$ of phase points with more than one singularities on their orbit we obtain the validity of the induction hypotheses (H1)---(H2) (and therefore (H3)---(H4), as well) for the considered cylindric billiard flow $\flow$ with $k$ cylinders. This finishes the inductive proof of the Theorem. \qed \bigskip \bigskip \Refs \widestnumber\key{B-Ch-Sz-T(1992)-II} \ref\key B(1979) \by L. A. Bunimovich \paper On the Ergodic Properties of Nowhere Dispersing Billiards \jour Commun. Math. Phys. \vol 65 \pages 295-312 \endref \ref\key B-Ch-Sz-T(2001) \by P. B\'alint, N. I. Chernov, D. Sz\'asz, I. P. T\'oth \paper Multi\-dimensional Semi\-dispersing Billiards: Singularities and the Fundamental Theorem \jour Manuscript \vol 2001 \endref \ref\key B-F-K(1998) \by D. Burago, S. Ferleger, A. Kononenko \paper A geometric approach to semi\-dispersing billiards \jour Ergod. Th. \& Dynam. Sys. \vol 18 \year 1998 \pages 303-319 \endref \ref\key Ch(1994) \by N. I. Chernov \paper Statistical Properties of the Periodic Lorentz Gas. Multidimensional Case \jour Journal of Statistical Physics \vol 74, Nos. 1/2 \year 1994 \pages 11-54 \endref \ref\key C-H(1996) \by N. I. Chernov, C. Haskell \paper Non-uniformly hyperbolic K-systems are \newline Bernoulli \jour Ergod. Th. \& Dynam. Sys. \vol 16 \year 1996 \pages 19-44 \endref \ref\key E(1978) \by R. Engelking \paper Dimension Theory \jour North Holland \year 1978 \endref \ref\key G(1981) \by G. Galperin \paper On systems of locally interacting and repelling particles moving in space \jour Trudy MMO \vol 43 \year 1981 \pages 142-196 \endref \ref\key He(1939) \by G. A. Hedlund \paper The Dynamics of Geodesic Flows \jour Bull. Amer. Math. Soc. \vol 45 \pages 241-260 \endref \ref\key Ho(1939) \by E. Hopf \paper Sta\-tis\-tik der ge\-o\-de\-ti\-schen Li\-ni\-en in Man\-nig\-fal\-tig\-kei\-ten ne\-ga\-tiver \newline Kr\"um\-mung \jour Ber. Verh. S\"achs. Akad. Wiss. Leipzig \vol 91 \pages 261-304 \endref \ref\key K-B(1994) \by A. Katok, K. Burns \paper Infinitesimal Lyapunov functions, invariant cone families and stochastic properties of smooth dynamical systems \jour Ergodic Theory Dyn. Syst. \vol 14, No. 4 \year 1994 \pages 757-785 \endref \ref\key K-S(1986) \by A. Katok, J.-M. Strelcyn \paper Invariant Manifolds, Entropy and Billiards; \newline Smooth Maps with Singularities \jour Lecture Notes in Mathematics \vol 1222 \newline \pages Springer Verlag \endref \ref\key K-S-Sz(1989) \by A. Kr\'amli, N. Sim\'anyi, D. Sz\'asz \paper Ergodic Properties of Semi--Dispersing Billiards I. Two Cylindric Scatterers in the 3--D Torus \jour Nonlinearity \vol 2 \pages 311--326 \endref \ref\key K-S-Sz(1990) \by A. Kr\'amli, N. Sim\'anyi, D. Sz\'asz \paper A ``Transversal'' Fundamental Theorem for Semi-Dis\-pers\-ing Billiards \jour Commun. Math. Phys. \vol 129 \pages 535--560 \endref \ref\key K-S-Sz(1991) \by A. Kr\'amli, N. Sim\'anyi, D. Sz\'asz \paper The K--Property of Three Billiard Balls \jour Annals of Mathematics \vol 133 \pages 37--72 \endref \ref\key K-S-Sz(1992) \by A. Kr\'amli, N. Sim\'anyi, D. Sz\'asz \paper The K--Property of Four Billiard Balls \jour Commun. Math. Phys. \vol 144 \pages 107-148 \endref \ref\key L-W(1995) \by C. Liverani, M. Wojtkowski \paper Ergodicity in Hamiltonian systems \jour Dynamics Reported \vol 4 \pages 130-202, arXiv:math.DS/9210229. \endref \ref\key O-W(1998) \by D. Ornstein, B. Weiss \paper On the Bernoulli Nature of Systems with Some Hyperbolic Structure \jour Ergod. Th. \& Dynam. Sys. \vol 18 \year 1998 \pages 441-456 \endref \ref\key P(1977) \by Ya. Pesin \paper Characteristic Exponents and Smooth Ergodic Theory \jour Russian Math. surveys \vol 32 \pages 55-114 \endref \ref\key Sim(1992-A) \by N. Sim\'anyi \paper The K-property of $N$ billiard balls I \jour Invent. Math. \vol 108 \year 1992 \pages 521-548 \endref \ref\key Sim(1992-B) \by N. Sim\'anyi \paper The K-property of $N$ billiard balls II. Computation of neutral linear spaces \jour Invent. Math. \vol 110 \year 1992 \pages 151-172 \endref \ref\key Sim(1999) \by N. Sim\'anyi \paper Ergodicity of hard spheres in a box \jour Ergod. Th. \& Dynam. Sys. \vol 19 \pages 741-766 \endref \ref\key Sim(2001) \by N. Sim\'anyi \paper Proof of the Boltzmann--Sinai Ergodic Hypothesis for Typical Hard Disk Systems \jour Manuscript \year 2001 \endref \ref\key Sim(2002) \by N. Sim\'anyi \paper The Complete Hyperbolicity of Cylindric Billiards \jour Ergod. Th. \& Dynam. Sys. \vol 22 \year 2002 \pages 281-302, arXiv:math.DS/9906139 \endref \ref\key S-Sz(1994) \by N. Sim\'anyi, D. Sz\'asz \paper The K-property of 4-D Billiards with Non-Orthogonal Cylindric Scatterers \jour J. Stat. Phys. \vol 76, Nos. 1/2 \pages 587-604 \endref \ref\key S-Sz(1999) \by N. Sim\'anyi, D. Sz\'asz \paper Hard Ball Systems Are Completely Hyperbolic \jour Annals of Math. \vol 149 \pages 35-96, arXiv:math.DS/9704229. \endref \ref\key S-Sz(2000) \by N. Sim\'anyi, D. Sz\'asz \paper Non-integrability of Cylindric Billiards and Transitive Lie Group Actions \jour Ergod. Th. \& Dynam. Sys. \vol 20 \pages 593-610 \endref \ref\key Sin(1963) \by Ya. G. Sinai \paper On the Foundation of the Ergodic Hypothesis for a Dynamical System of Statistical Mechanics \jour Soviet Math. Dokl. \vol 4 \pages 1818-1822 \endref \ref\key Sin(1970) \by Ya. G. Sinai \paper Dynamical Systems with Elastic Reflections \jour Russian Math. Surveys \vol 25:2 \year 1970 \pages 137-189 \endref \ref\key Sin(1979) \by Ya. G. Sinai \paper Development of Krylov's ideas. Afterword to N. S. Krylov's ``Works on the foundations of statistical physics'', see reference [K(1979)] \jour Princeton University Press \year 1979 \endref \ref\key S-Ch(1982) \by Ya. G. Sinai, N.I. Chernov \paper Entropy of a gas of hard spheres with respect to the group of space-time shifts \jour Trudy Sem. Petrovsk. \vol No. 8 \year 1982 \pages 218-238 \endref \ref\key S-Ch(1987) \by Ya. G. Sinai, N.I. Chernov \paper Ergodic properties of certain systems of 2--D discs and 3--D balls \jour Russian Math. Surveys \vol (3) 42 \year 1987 \pages 181-207 \endref \ref\key Sz(1993) \by D. Sz\'asz \paper Ergodicity of classical billiard balls \jour Physica A \vol 194 \pages 86-92 \endref \ref\key Sz(1994) \by D. Sz\'asz \paper The K-property of `Orthogonal' Cylindric Billiards \jour Commun. Math. Phys. \vol 160 \pages 581-597 \endref \ref\key Sz(2000) \by D. Sz\'asz \paper Ball-avoiding theorems \jour Ergod. Th. \& Dynam. Sys. \vol 20 \year 2000 \pages 1821-1849 \endref \ref\key V(1979) \by L. N. Vaserstein \paper On Systems of Particles with Finite Range and/or Repulsive Interactions \jour Commun. Math. Phys. \vol 69 \year 1979 \pages 31-56 \endref \ref\key W(1985) \by M. Wojtkowski \paper Invariant families of cones and Lyapunov exponents \jour Ergod. Th. \& Dynam. Sys. \vol 5 \pages 145-161 \endref \ref\key W(1986) \by M. Wojtkowski \paper Principles for the Design of Billiards with Nonvanishing Lyapunov Exponents \jour Commun. Math. Phys. \vol 105 \pages 391-414 \endref \endRefs \bye ---------------0207241512621--