math0410464/4qp.tex
1: \documentclass[12pt]{amsart}
2: \usepackage{amsmath,amssymb,amsthm,epsfig}
3: \textheight=22cm
4: \textwidth=14cm
5: \voffset=-1cm
6: \hoffset=-1cm
7: \def\torus{\mathbb T}
8: \def\real{\mathbb R}
9: \def\complex{\mathbb C}
10: \def\integer{\mathbb Z}
11: \def\const{\text{const}}
12: \def\rk{\qopname\relax o{rk}}
13: \def\Im{\qopname\relax o{Im}}
14: \def\Iso{\qopname\relax o{Iso}}
15: \newtheorem{theorem}{Theorem}
16: \newtheorem{lemma}{Lemma}
17: \theoremstyle{remark}
18: \newtheorem{ex}{Example}
19: \newtheorem{case}{Case}
20: \newtheorem{conjecture}{Conjecture}
21: \newtheorem{remark}{Remark}
22: \author{I.\,Dynnikov}
23: \author{ S.\,Novikov}
24: \address{Dept.\ of Mech.\ and Math., Moscow State University,
25: Moscow 119992 GSP-2, Russia}
26: \email{dynnikov@mech.math.msu.su}
27: \address{Landau Institute for theoretical physics, Kosygina str.~2,
28: Moscow 119334, Russia; IPST, University of Maryland, College Park,
29: MD 20742, USA}
30: \email{novikov@glue.umd.edu}
31: \title{Topology of quasiperiodic functions on the plane}
32: \thanks{The work of I.\,Dynnikov was supported in part by
33: Russian Foundation for Basic Research (grant no.~02-01-00659);
34: the work of S.\,Novikov was supported in part by
35: the Council of the Russian Academy of Science (grant ``Mathematical
36: methods of nonlinear dynamics'')}
37: 
38: 
39: 
40: \begin{document}
41: 
42: \maketitle
43: 
44: 
45: \begin{abstract}This article describes a topological theory of
46: quasiperiodic functions on the plane. The development of
47: this theory was started (in different
48: terminology) by the Moscow topology group in early 1980s.
49: It was motivated by the needs of solid state physics, as a partial
50: (nongeneric) case of Hamiltonian foliations of  Fermi
51: surfaces with multivalued Hamiltonian function~\cite{N}. The
52: unexpected discoveries of their topological properties that were made in
53: 1980s \cite{Z,N1} and 1990s \cite{D,D',D1} have finally led to
54: nontrivial physical conclusions \cite{NM,NM1} along the lines of
55: the so-called geometric strong magnetic field limit
56: \cite{LAK}. A very fruitful new point of view comes from
57: the reformulation of that problem
58: in terms of quasiperiodic functions and an extension to higher dimensions
59: made in 1999 \cite{N2}. One may say that,
60: for single crystal normal metals put in a magnetic field,
61: the semiclassical trajectories of electrons in the space of
62: quasimomenta are exactly the level lines of the quasiperiodic function with
63: three quasiperiods that is the dispersion relation restricted to a
64: plane orthogonal to the magnetic field. General studies of the
65: topological properties of levels of quasiperiodic functions on the
66: plane with any number of quasiperiods were started in 1999 when
67: certain ideas were formulated for the case of four quasiperiods
68: \cite{N2}. The last section of this work contains a complete
69: proof of these results based on the technique developed in \cite{D2,D3}.
70: Some new physical applications of the general problem were found
71: recently \cite{M}.
72: \end{abstract}
73: 
74: \section{Quasiperiodic functions}
75: 
76: Let $\torus^n=\real^n/\integer^n$ denote the
77: $n$-dimensional torus,
78: $\nu\!:\real^n\rightarrow\torus^n=\real^n/\integer^n$ the standard
79: projection.
80: 
81: We say that a real smooth function $\varphi(y)=\varphi(y^1,\dots,y^k)$
82: on the $k$-plane $\real^k$ is \emph{quasiperiodic} with $n$
83: quasiperiods (frequencies) if it can be represented in the form
84: $\varphi(y)=f(x(y))$:
85: \begin{equation}
86: \varphi=f\circ\nu\circ \iota,
87: \end{equation}
88: where $\iota:\real^k\rightarrow\real^n$ is an affine imbedding:
89: $$x^s=a^s_ry^r+x^s_0,$$
90: $f=f(x):\torus^n\rightarrow\real$ is some
91: smooth function, and $n\geqslant 2$ is the minimal possible integer for
92: which such a function $f$ and an affine imbedding $\iota$ exist. Here
93: $s=1,\dots,n$ and $r=1,\dots,k$. In the theory of quasicrystals, people call
94: the space $\real^k$ (where $k=2,3$) the
95: \emph{physical space} and the space $\real^n$
96: the \emph{superspace}. Every $n$-periodic function $f(x)$ generates a
97: family of \emph{descendants}, which are obtained by varying
98: the initial vector
99: $x_0=(x^1_0,\dots,x^n_0)$ in the superspace $\real^n$.
100: Any two descendants $\varphi_1(y)$, $\varphi_2(y)$
101: of the same function $f$ are said to be
102: \emph{related}. They have the same frequencies
103: and obtain the following property: for any $\varepsilon>0$,
104: there is a shift $y\mapsto y+a$ in the physical space
105: such that the shifted function
106: $\varphi_2(y+a)$ is $\varepsilon$-close to $\varphi_1(y)$:
107: $$|\varphi_2(y+a)-\varphi_1(y)|<\varepsilon\qquad\forall y\in\real^k.$$
108: 
109:  Any
110: linear function $\lambda:\real^k\rightarrow\real$ of the form
111: $\lambda(y)=\ell(x(y))$ or $\lambda=\ell\circ \iota$, where the
112: linear function $\ell:\real^n\rightarrow\real$ belongs to the dual
113: (or ``reciprocal'') lattice $(\integer^n)^*$ (\emph{i.e.}, we have
114: $\ell(\integer^n)\subset\integer$) is called a \emph{frequency} of
115: $\varphi$. The set of all frequencies form a free abelian group
116: with $n$ natural  generators
117: $\lambda_1=\ell^1\circ\iota,\dots,\lambda_1=\ell^n\circ\iota$ where
118: the functions
119: $\ell^s(x)=x^s$, $s=1,\ldots,n$, are dual to
120: the basic periods. We call this group
121: the \emph{group of frequencies}. It is a free
122: abelian subgroup $\Gamma^*$ of the dual vector space $\real^*$,
123: and it is the
124: same for the whole family of related quasiperiodic functions (descendants of the
125: same $n$-periodic function $f$).
126: 
127: Analytically, any $n$-periodic function can be presented in the form
128: of a trigonometric series
129: 
130: $$f(x)=\sum_{\ell\in(\integer^n)^*} c_\ell\exp\big(2\pi
131: i\ell(x)\big)$$
132: 
133: Therefore, any  quasiperiodic function can be presented in a
134: similar form:
135: $$\varphi(y)=\sum_{\lambda\in\Gamma^*}b_{\lambda}\exp\big(2\pi
136: i\lambda(y)\big)= \sum_mb_m\exp\Big(2\pi i
137: \sum_{s=1}^nm_s\lambda_s(y)\Big),$$
138: where $m=(m_1,\dots,m_n)
139: \in\integer^n$. By definition, the set of basic frequencies
140: $\lambda_s$ generates the space $\real^k$ over the field $\real$.
141: 
142: For the space $\real^k$ endowed with a Euclidean metric there is a natural
143: identification $\real^k\cong(\real^k)^*$, so the subgroup of
144: frequencies can be treated as a subgroup $\Gamma\cong\Gamma^*\subset
145: \real^k$ in the physical space $\real^k$.
146: 
147: There is an affine symmetry semigroup associated with each family
148: of related quasiperiodic functions. By definition, this semigroup
149: $\widetilde{G}$ consists of all affine transformations
150: $$g:\real^k\rightarrow \real^k$$
151: of the physical space
152: such that
153: $$g(\Gamma^*)\subset\Gamma^*,$$
154: where $\Gamma^*$ is treated as
155: a subset of the group of translations: $\Gamma^*\subset \real^k$.
156: For the Euclidean space $\real^k$ we define the \emph{symmetry group}
157: $G\subset\widetilde{G}$ consisting only of isometries $g$ such
158: that $g(\Gamma^*)=\Gamma^*$.
159: 
160: This group satisfies the general definition of a
161: quasicrystallographic group introduced by S.\,Novikov and
162: A.\,Veselov in 1980s in order to answer the question: what is the
163: symmetry of quasicrystals (see~\cite{Pi})? According to that
164: definition the intersection $G\bigcap\real^k\subset\Iso(\real^k)$
165: of a quasicrystallographic group
166: with the subgroup of translations should be a finitely generated free
167: abelian group. In our case it is exactly the group $\Gamma^*$. The
168: definition allows  the ``rotational'' quotient group
169: $G/(G\bigcap \real^k)\subset O_k$ to be infinite.
170: S.\,Piunikhin studied these
171: groups for $k=2,3$ in a series of works (see \cite{Pi}).
172: 
173: \begin{ex} Consider the two-dimensional case, $k=2$. Let $\theta$ be a
174: unimodular complex number, $|\theta |=1$, $\theta=\exp(i\psi)$,
175: satisfying the equation
176: $$P(\theta)=\theta^n+a_1\theta^{n-1}+\ldots+a_{n-1}\theta+1=0,$$
177: where all coefficients are  integer-valued, $a_s\in
178: \integer$, and we have $a_s=a_{n-s}$. The complex numbers
179: (or real two-vectors)
180: $\lambda_1=1, \lambda_2=\theta,\dots, \lambda_n=\theta^{n-1}$
181: generate a group of frequencies $\Gamma^*\subset
182: \complex=\real^2$  with nontrivial rotational symmetry $g\rightarrow
183: g\exp(i\psi)$. It is easy to find such a polynomial $P$ with
184: a root at $\theta=\exp(i\psi)$, where the
185: ratio $\psi/2\pi$ is irrational. There are very complicated
186: quasicrystallographic groups for $k=3$ (see \cite{Pi}).
187: 
188: \end{ex}
189: 
190: \section{Quasiperiodic functions in analysis, geometry, and
191: dynamical systems motivated by natural sciences}
192: 
193: \subsection{Quasiperiodic functions on the real line}
194: 
195: Consider the case $k=1$. In the XIX century,
196: one-dimensional quasiperiodic functions
197: with $n$ quasiperiods appeared in the theory of
198: completely integrable Hamiltonian systems of the classical
199: mechanics with $n$ degrees of freedom. According to so-called
200: Liouville's theorem, the integrability follows from the existence of
201: $n$ smooth independent pairwise commuting integrals of motion. If their
202: common level sets are compact, then the time dependence of the space
203: coordinates along a trajectory can be described by
204: quasiperiodic functions $x^r(t)$ with (at most) $n$ quasiperiods.
205: So, all studies of perturbations of
206: completely integrable systems should start with quasiperiodic
207: unperturbed background. A lot of fundamental work has been done in
208: this area (see \cite{DKN}).
209: 
210: \subsection{Quasiperiodic functions in the theory of nonlinear PDE}
211: Completely integrable PDE systems of the theory of solitons give rise
212: to quasiperiodic functions with $k>1$. There are very famous
213: ($1+1$) PDE systems such as KdV ($u_t=6uu_x+u_{xxx}$) or sine-Gordon
214: ($u_{tt}=u_{xx}+\sin \{u(x,t)\}$), which are completely integrable by the
215: so-called \emph{inverse scattering transform method for rapidly
216: decreasing initial values}. A countable number of continuous
217: families of exact
218: smooth real ``finite-gap'' solutions of these equations were discovered
219: in 1970s (see~\cite{DKN}). These solutions are quasiperiodic
220: functions in $x,t$, and depend on many parameters $a$, $a'$:
221: $$u(x,t)=F(xU+Vt+U_0;a)$$
222: for KdV, and
223: $$\exp\big(iu(x,t)\big)=F'(U'x+V't+U'_0;a')$$ for sine-Gordon. Here
224: $u(x,t)$ is real in both cases, $F,F'$ are $n$-periodic smooth
225: functions in $n$ variables (\emph{i.e.}, smooth functions on the real
226: $n$-torus). They can be expressed through special functions, namely,
227: theta-functions of a hyperelliptic Riemann surface
228: of genus~$n$. $U$, $U'$, $V$, $V'$ are the $n$-vectors of
229: periods of some Abelian differentials of the second kind
230: (see \cite{N1}). Let us
231: mention that, for the sine-Gordon system, the function $u=1/i\,\log
232: F'$ is generically a multivalued function on the ``real''
233: $n$-torus imbedded in the complex $2n$-dimensional Jacoby torus
234: associated with a complex hyperelliptic Riemann surface. Here we
235: have $k=2$. For famous completely integrable $(2+1)$ systems
236: (like KP, and others) one comes to quasiperiodic solutions of the form
237: $u(x,y,t)$, which are quasiperiodic functions
238: in $k=3$ physical variables. When studying the
239: dependence of the solution on so-called higher times one may arrive at any
240: value of $k$.
241: 
242: \subsection{Quasiperiodic functions and quasicrystals}
243: 
244: Completely different examples come from solid
245: state physics. In 1980s a new type of 2D and 3D media was
246: discovered. People named them ``quasicrystals''. The optical
247: analysis of the location of atoms gave an evidence for the group
248: of frequencies being incompatible with an ordinary crystal structure.
249: For example,
250: for $k=2$, the observed group of frequencies $\Gamma^*$ might be generated
251: by the 5th roots of unity:
252: $$\lambda_r=\eta^r\in\Gamma^*,\quad r=0,1,2,3,\quad\eta^5=1,\quad P(\eta)=0,$$
253: where
254: $$P(z)=z^4+z^3+z^2+z+1.$$
255: 
256: Recall that our extension of the idea of symmetry allows the
257: rotational symmetry to be even infinite.
258: 
259: There are two mathematical models of quasicrystals.
260: Let us think of atoms in
261: the physical space $\real^k$ as being located in a discrete set of
262: points $x_A$ such that
263: there exists a couple of positive ``radii'' $\rho_1$, $\rho_2$ with the properties:
264: \begin{itemize}
265: \item[a.]
266: We have $|x_A-x_{A'}|\geq\rho_2$ for all pairs $A,A'$ with $A\neq
267: A'$;
268: \item[b.]
269: For every point $x\in \real^k$, there exists a point $x_{A}$
270: such that $|x_A-x|\leqslant \rho_1$.
271: \end{itemize}
272: We call this set of points
273: \emph{quasiperiodic} if the distribution $\sum_A\delta(x-x_A)$ can be
274: decomposed into a Fourier series with finitely generated free
275: abelian group of frequencies $\Gamma^*$.
276: 
277: In another model, our
278: physical space $\real^k$ is endowed with
279: a ``quasiperiodic tiling''. This means the following:
280: \begin{itemize}
281: \item[a.] The space is
282: covered by countably many polytopes $P_B$,
283: $\real^k=\bigcup_B P_B$, where $P_B\bigcap P_{B'}$ is a face for any pair $B,B'$.
284: \item[b.]
285: Up to shift, there is only a finite number of different polytopes
286: $P_1,\dots,P_q$ among them.
287: \item[c.]
288: Let us associate some constant $c_q$ with every polytope $P_q$
289: and consider a function that is equal to $c_j$ everywhere
290: in the interior of any $P_B$ obtained from $P_j$ by a shift. We obtain a
291: piecewise constant function $c(x)$ in $x\in \real^k$ defined
292: (at a full measure set)
293: by our tiling and the choice of the constants $c_j$.
294: The tiling is said to be \emph{quasiperiodic} if, for every
295: choice of constants $c_j$, the function $c(x)$ is quasiperiodic,
296: \emph{i.e.}, can be presented in the form of a
297: trigonometric series with finitely generated free abelian group $\Gamma^*$ of
298: frequencies.
299: \end{itemize}
300: 
301: 
302: There is a famous tiling of
303: the plane $\real^2$ by rhombi of two types: one with
304: angles $\pi/5$ and $4\pi/5$,
305: and the other with angles $2\pi/5$ and $3\pi/5$.
306: It is called the Penrose tiling.
307: This tiling is quasiperiodic, which was discovered a few years later
308: after Penrose's original work
309: (see the history and details of the subject in~\cite{Pi}). An
310: interesting idea of ``local rules'' was developed by
311: physicists and mathematicians in order to explain the growth of
312: quasicrystals in terms of tilings. In this model, the atoms are
313: located at the vertices of the tiles.
314: 
315: Both models can be obtained from the following construction. Let a
316: ``superlattice'' $\Gamma$ of full rank be given
317: in the superspace $\real^n$, and the
318: superspace be presented as the direct sum
319: $\real^n=\real^k\bigoplus\real^{n-k}$,
320: where $\real^k$ is the physical subspace.
321: Let
322: $p:\real^n\rightarrow \real^k$ and $q:\real^n\rightarrow
323: \real^{n-k}$ be the natural projections.
324: Fix a
325: finite $(n-k)$-polyhedron $D\subset
326: \real^{n-k}$ and consider the ``tubular $D$-neighborhood''
327: $D_q=q^{-1}(D)\subset \real^n$ of the physical subspace
328: $\real^k\subset\real^n$.
329: Assume that the boundary of the polyhedron $D$ is disjoint from
330: $q(\Gamma)$, or equivalently, $\partial D_q\cap\Gamma=\varnothing$.
331: Then the set of points
332: $$p(\Gamma\cap D_q)\subset\real^k$$
333: is quasiperiodic in the sense of the definition given above.
334: 
335: By taking a certain polytope
336: decomposition of the space $\real^n$ associated with the lattice
337: $\Gamma$ and the polyhedron $D$, one obtains a quasiperiodic
338: tiling of $\real^k$
339: whose tiles are the intersections of $\real^k$ with
340: the $n$-cells of the decomposition (see survey
341: article \cite{Pi}).
342: 
343: Very interesting examples of nontrivial
344: symmetry groups come from the superspace $\real^4$ endowed with
345: the Minkovski metric and
346: a lattice $\Gamma\cong\integer^4$ such that the
347: physical subspace $\real^2$, which is spacelike
348: (\emph{i.e.}\ Euclidean), is
349: invariant under some lattice-preserving mapping from the group
350: $O(3,1)$.
351: 
352: The superspaces $\real^{l,m}$, where $l+m=n$, might also appear in
353: interesting cases.
354: 
355: \subsection{Quasiperiodic functions in the theory of conductivity}
356: 
357: Here we describe the situation that is the main motivation for
358: our topological and dynamical theory.
359: 
360: For every single
361: crystal normal metal, we have a lattice $\Gamma$ in the physical
362: space $\real^3$. However, our geometrical constructions will live in
363: a completely different space, namely the $3$-torus of
364: quasimomenta $\torus^3$, which
365: is the quotient space of the \emph{dual} $3$-space $(\real^3)^*\cong\real^3$
366: by the dual
367: (reciprocal) lattice $\Gamma^*\cong\integer^3$. The ``Bloch'' states of quantum
368: electrons are parameterized by pairs $(m,p)$,
369: where $p$ is a point in the space of
370: quasimomenta, $p\in \torus^3=\real^3/\Gamma^*$, and $m$ is a
371: natural number, which is the index of
372: a branch of the dispersion relation
373: $f(p)=\epsilon_m(p):\torus^3\rightarrow \real$. In what follows
374: we will always deal with just one branch only, so we
375: drop the index $m$ in the notation. We assume that $f(p)=\epsilon(p)$ is a Morse
376: function on the $3$-torus or, in other words,
377: a three-periodic Morse function on the
378: covering Euclidean space $\real^3$. At zero temperature all
379: electrons occupy the ``Dirac sea'' $\epsilon(p)\leqslant \epsilon_F$
380: where the ``Fermi energy'' $\epsilon_F$ depends on the number of free
381: electrons in the metal. We assume that $\epsilon_F$ is a regular value
382: for the Morse function $f=\epsilon(p)$. At low temperatures we
383: are dealing only with
384: ``excited'' electrons nearby the
385: Fermi level $\epsilon(p)=\epsilon_F$.
386: 
387: The Fermi level looks geometrically as a
388: two-dimensional surface $M_F\subset \torus^3$
389: in the space of quasimomenta. This
390: surface is nonsingular and homologous to zero in the $3$-torus. Let
391: us assume that it is connected.
392: 
393: The \emph{topological rank} $r(M_F)$
394: of the Fermi surface is defined as the rank of the image
395: of the first homology group of $M_F$ under the mapping
396: $i_*:H_1(M_F,\integer)\rightarrow H_1(\torus^3,\integer)\cong\integer^3$
397: induced by the inclusion $i:M_F\hookrightarrow\torus^3$.
398: Since $i_*(H_1(M_F))$ is a sublattice in $\integer^3$,
399: we always have $r\in\{0,1,2,3\}$.
400: 
401: For
402: example, the topological rank of the Fermi surface of
403: lithium is equal to zero (in this case, the Fermi
404: surface looks like a topological $2$-sphere), whereas
405: it is equal to three for
406: copper, gold, platinum, and some other noble metals. For gold, for
407: example, the genus of the Fermi surface is equal to four.
408: 
409: The problem that we will consider is most difficult when the topological
410: rank of the Fermi surface is maximal possible, \emph{i.e.}, equal to three.
411: One can easily show that the genus of the Fermi surface must be
412: greater than or equal to the topological rank.
413: 
414: Interesting dynamical phenomena occur in the presence of a
415: magnetic field. In the semiclassical approximation, an electron,
416: which is considered as a point in the space of quasimomenta, moves
417: along constant energy lines in the plane $\real^2_{B,p_0}$
418: orthogonal to the magnetic field $B$ and passing through the
419: initial position $p_0$ of the quasimomentum.
420: 
421: One may say that this is a
422: Hamiltonian system on the $3$-torus of quasimomenta with Poisson
423: bracket $$\{p_j,p_l\}=\frac ec\,B_{jl}=\frac ec\,\varepsilon_{jlq}B^q$$ and
424: Hamiltonian $f=\epsilon(p)$:
425: $$\frac{dp_j}{dt}=\{p_j,\epsilon(p)\},$$
426: so
427: the motion preserves the energy and a linear Casimir of the Poisson
428: bracket. The level sets of this Casimir are planes orthogonal to the magnetic
429: field. The trajectories can be treated as  leaves of the
430: Hamiltonian foliation on the Fermi surface given by the
431: equation $\omega=0$ where $\omega$ is the following closed $1$-form:
432: $\omega=\sum_jB^jdp_j|_{M_F}$.
433: 
434: According
435: to the ``strong magnetic field limit'' principle worked out by
436: I.\,Lifshitz, M.\,Azbel, M.\,Kaganov, and V.\,Peschanski in early 1960s,
437: all essential properties of the electrical conductivity in the presence of
438: a reasonably strong uniform magnetic field $B$ should follow from the
439: structure of the
440: dynamical system on the Fermi surface described above (see
441: \cite{LAK,Ab,Ki,Zi}). For ordinary normal metals (like gold, for
442: example) one may use this approximation for magnetic fields strong
443: enough in the human sense (like
444: $1\,\mathrm{Tesla}<|B|<10^3\,\mathrm{Tesla}$ for low
445: temperatures; recall: $1\,\mathrm{Tesla}=
446: 10^4\,\mathrm{Gauss}$). If the magnetic field is too strong, then
447: the semiclassical approximation will not be valid.
448: If the magnetic field is too weak, then the electron quasimomentum drift will
449: be too slow, and the distance that the quasimomentum passes for the
450: characteristic time of the electron free motion will become
451: insufficient to affect the observable conductivity.
452: 
453: However, in 1960s the study of the just mentioned dynamical system
454: was only started. Some conceptual mistake
455: was then made in \cite{LP} and further investigation
456: was stopped, and resumed only many years later in works
457: \cite{N,Z,N1,D,D1,NM,NM1,D2,D3,DM,MN,MN1,MN2}.
458: 
459: What is crucial for us here is following:
460: 
461: \medskip
462: \centerline{\parbox{0.8\textwidth}{%
463: the electron trajectories coincide with connected components of
464: the level curves $\epsilon(p)=\epsilon_F$ of the function
465: $\epsilon$ restricted to the planes orthogonal
466: to the magnetic field $B$; in other words, they are connected
467: components of the level curves of functions that form a family of
468: related quasiperiodic functions with three quasiperiods. }}
469: \medskip
470: 
471: In work \cite{N2} an extension of these studies to
472: a larger number of quasiperiods was started. In particular, some
473: new  ideas and results   were formulated for the case $n=4$. The
474: present  work contains the first complete proof of those (properly corrected)
475: statements. The proof is based on the topological technique developed in
476: \cite{D2,D3}.
477: 
478: Let a constant Poisson bracket $B_{jk}$ of rank two be given
479: on the $n$-torus. Then every Hamiltonian
480: $f(p)=\epsilon(p):\torus^n\rightarrow \real$ defines a Hamiltonian
481: system whose trajectories are exactly the level lines
482: $\epsilon(p)=\const$ of the restriction
483: of the function $\epsilon$ to the planes $\real^2_{B,a}$
484: defined as follows.
485: There exist exactly $n-2$ independent linear
486: Casimirs $K_1,\dots,K_{n-2}$, $K_j(p)=K_j^lp_l$, such that
487: $\{p_s,K_j\}=0=K_j^lB_{sl}$. We put
488: $$\real^2_{B,a}=\{K_1=a_1,\dots,K_{n-2}=a_{n-2}\},$$
489: where $a=(a_1,\dots,a_{n-2})$.
490: So our trajectories
491: are exactly the levels of quasiperiodic
492: functions on the two-planes $\real^2_{B,a}$, which form
493: the family of descendants of the $n$-periodic function
494: $\epsilon(p)$. They depend on the constants $a_1,\dots,a_{n-2}$.
495: 
496: Topological study of this problem is the central part of this
497: article.
498: 
499: 
500: Modern experimental technology allows to construct surfaces with
501: a variety of prescribed small fluctuations. In particular, it is
502: possible to make  a quasiperiodic construction with any number of
503: quasiperiods. It presents us a two-dimensional weak quasiperiodic
504: electric potential $V(x,y)$. In a strong magnetic
505: field $B$ electrons move along the surface. After averaging we
506: obtain a slow motion along the level curves $V(x,y)=\const$. These
507: studies, experimental and theoretical, were done originally for
508: periodic potentials with $n\leqslant 2$ periods only (see
509: \cite{B}), but it was pointed out in work \cite{M} that
510: quasiperiodic potentials can also appear here; new predictions
511: were made for quasiperiodic cases with three and four quasiperiods
512: based on the topological results obtained in a series of works
513: of the present authors (\cite{N2,D3,MN2}).
514: 
515: \section{Topology and dynamics of quasiperiodic functions on the
516: plane: the case of three quasiperiods. The electrical conductivity in
517: metals}
518: 
519: We address the following general question. How may the level lines
520: $\varphi=\const$ of a quasiperiodic function $\varphi$ on the plane
521: with $n$
522: quasiperiods look like? In a generic situation, such a level line
523: is a one-dimensional submanifold of $\real^2$, i.e.\ a union of
524: curves. We will call these curves ``trajectories'' because in our studies they
525: have been appearing as semiclassical electron trajectories
526: on the Fermi surface in the presence of a magnetic field since early 1980s when
527: this problem was posed in work \cite{N} as a problem of
528: topology and dynamical systems. It corresponds to the case of three
529: quasiperiods only. Some of those curves may be closed in $\real^2$
530: (compact) and others nonclosed in $\real^2$ (open). Let us ask the
531: following questions.
532: 
533: \begin{description}
534: \item[Question 1] Is the size of the compact trajectories uniformly bounded (for a
535: fixed level of $\varphi$)?
536: \item[Question 2] Do the open trajectories have some nice asymptotic behavior?
537: \end{description}
538: 
539: The first results were
540: obtained in work \cite{Z}. It became clear in the second half
541: of 1980s  that the proper form of Question~2 is the following:
542: does any open trajectory has a ``strong asymptotic direction'' in
543: $\real^2$, \emph{i.e.}, lie in a strip of uniformly bounded width
544: and passes through the strip
545: ``from $-\infty$ to $+\infty$''? This specification of the problem
546: was made in article \cite{N1}. In work \cite{D} the
547: results of \cite{Z} were improved accordingly to the new formulation of
548: the problem. The important breakthrough was made in work
549: \cite{D1}, but for a long time there was no applications.
550: Physical applications were found later in works~\cite{NM,NM1}.
551: 
552: In the physically important  case $n=3$, the positive answer to
553: our Question~1 follows easily from a quite elementary argument.
554: For $n>3$ it is more difficult, and it will be discussed later.
555: Question~2 is highly nontrivial already for $n=3$.
556: As mentioned above, the asymptotic behavior of open electron
557: trajectories, \emph{i.e.}, of open connected components
558: of a level line of a quasiperiodic
559: function with three quasiperiods, was studied in
560: \cite{Z,D,D1}. It became finally
561: clear after work \cite{D1} that for the family of
562: related quasiperiodic function corresponding to
563: a ``typical'' direction of the
564: magnetic field (which is regarded as the direction of a
565: plane $\real^2\subset \real^3$),
566: either their level lines do not have open components at all or
567: the open components all have a strong asymptotic direction.
568:  The
569: latter means that each open curve has a parametrization
570: $\gamma(t)$ such that the following holds for some nonzero two-vector
571: $(x_1,y_1)$:
572: \begin{equation}\label{strong}
573: \gamma(t)=(x_0,y_0)+t\cdot(x_1,y_1)+O(1).\end{equation}
574: Below we will explain precisely
575: what  `a typical direction' means,
576: and provide references to the papers containing proofs of
577: the corresponding results. In the first work \cite{Z}
578: completed in the note
579: \cite{D}, this type of result was obtained for the special case of
580: small perturbations of a magnetic field having ``rational''
581: direction. We shall return to this special case in the next section
582: where we discuss the
583: quasiperiodic functions with four quasiperiods.
584: 
585: Applications of this studies to explaining the electrical
586: conductivity in a strong magnetic field are presented in
587: works \cite{NM,NM1}. They are based on the results of the Lifshitz
588: school of 1960s. Physicists calculated the contribution of
589: individual trajectories of simple types to the conductivity
590: tensor. These calculations have become a part of textbooks (see
591: \cite{LAK,Ab}). In the case when all trajectories are
592: compact the conductivity
593: components orthogonal to the magnetic field $B$ decrease as $|B|^{-1}$ or
594: $|B|^{-2}$ when $|B|$ grows while the direction of $B$
595: remains fixed. Some special examples of open trajectories lying in
596: finitely wide strips were found at the same time and their
597: contribution to the conductivity was
598: calculated. As pointed out in \cite{NM,NM1},
599: one can easily extend the just mentioned
600: calculation to the case of general trajectories of the same type.
601: The projection to the plane orthogonal to $B$
602: of the part of the conductivity tensor contributed by such a trajectory
603: has two eigenvalues one of which is zero and the other nonzero.
604: Since the contribution of closed trajectories tends to zero
605: when $|B|$ grows, the observable conductivity tensor
606: for a strong enough $B$ will depend on open trajectories only.
607: 
608: \smallskip
609: \centerline{\parbox{0.8\textwidth}
610: {However, the observable physical
611: conductivity tensor is formed by the contributions of all electron
612: trajectories as the sum of them. What conclusion about this tensor
613: can be made from the qualitative dynamical properties of
614: that system, which was defined on the quantum level?}}
615: \smallskip
616: 
617: In order to obtain a nontrivial new physical result, one needs
618: more than the theorems explicitly formulated in~\cite{D1}.
619: But luckily an additional crucial property also holds
620: for our dynamical system,
621: and this can be extracted from the proofs of the main theorems
622: of works \cite{Z,D1}. The property, which we call
623: \emph{topological resonance}, implies the following for the behavior
624: of trajectories.
625: 
626: \smallskip
627: \centerline{\parbox{0.8\textwidth}{For a ``typical'' family $\{\varphi_a\}$
628: of related quasiperiodic functions $\varphi_a(p)=\epsilon(p)|_{\real^2_{B,a}}$,
629: the strong asymptotical direction $\eta_B$ of noncompact trajectories is the
630: same for all trajectories.
631: Moreover, there exist an integral two-plane $\mu\subset
632: \real^3$ (\emph{i.e.}, a plane generated by two reciprocal lattice vectors,
633: $\mu\cap\integer^3\cong\integer^2$)
634: such that $\eta_B$ has the direction of the intersection of $\mu$
635: with the plane orthogonal to the magnetic field:
636: $$\eta_B\in\mu\cap \real^2_B.$$
637: This integer plane $\mu$ is
638: locally rigid, \emph{i.e.}, it remains unchanged under small
639: variations of the direction of the magnetic field.}}
640: \smallskip
641: 
642: The topological resonance property of our dynamical system
643: makes possible serious applications. It was missed in the classical works
644: of physicists, and a conceptual mistake was made in \cite{LP},
645: where calculations led to a result contradicting to this
646: property. This mistake was revealed and corrected only in works
647: \cite{NM,NM1,MN}.
648: 
649: For a strong enough $B$, the direction $\eta_B$ is a
650: zero eigenvector of the projection of the conductivity tensor to the
651: plane orthogonal to the magnetic field. The
652: integral plane $\Pi\subset \real^3$ is
653: directly observable by measuring the zero eigenvector $\eta_B$ for two
654: or more magnetic fields $B$ close to each other.
655: 
656: We refer the reader to recent article \cite{MN} for a more detailed
657: physical discussions.
658: 
659: Let us describe the picture topologically. Consider all our objects
660: in the universal covering space $\real^3$ with the reciprocal lattice
661: $\Gamma^*=\integer^3\subset \real^3$ and the three-periodic Fermi surface
662: $$\widehat{M_F}=\nu^{-1}(M_F)\subset \real^3$$
663: covering the compact one
664: $$M_F\subset
665: \torus^3.$$
666: (Recall that $\nu$ stays for the standard projection $\real^3\rightarrow\torus^3$.)
667: The electron trajectories in the covering space are
668: connected components of the intersections of the
669: three-periodic Fermi surface with planes orthogonal to $B$. Let
670: $M_0(B)$ be the closure of the union of all compact trajectories,
671: and let $L(B)$ be the closure of its complement in the Fermi surface:
672: $$L(B)=\overline{M_F\setminus M_0(B)}.$$
673: 
674: Let $L_l(B)$ be the connected components of $L(B)$.
675: In the typical case, $L(B)$ is a compact two-manifold with boundary
676: $$\partial
677: L(B)=\bigcup_{l,s} \beta_{ls},$$
678: where
679: $$\partial L_l(B)=\bigcup_s\beta_{sl}.$$
680: All boundary  curves $\beta_{sl}$ are
681: saddle connection cycles. In the typical case, we may assume that
682: every cycle $\beta_{sl}$ joins a saddle critical point to itself, since
683: all the other cases have
684: positive codimension in the appropriate functional space.
685: (In particular, we assume that there
686: is no rational linear dependence between the components of $B$,
687: and that the Hamiltonian foliation defined by $\omega=\sum
688: B^idp_i|_{M_F}=0$ has only Morse singularities and
689: does not have saddle connections between different saddles.)
690: 
691: The part $M_0$ of the Fermi surface
692: can be presented as the union of ``cylinders'' $Z_q$,
693: $M_0=\bigcup_q Z_q$,
694: whose interior consists of regular compact trajectories and
695: ``bases'' are either saddle connection cycles or
696: isolated points (centers). There are finitely many such
697: cylinders, and they are obviously compact.
698: This immediately implies a positive
699: answer to Question~1 posed in the beginning of this section:
700: 
701: \begin{center}the size of all compact trajectories is uniformly bounded.
702: \end{center}
703: 
704: 
705: We call the
706: pieces $L_l$ of the Fermi surface the \emph{carriers of open
707: trajectories}. By construction, every open trajectory
708: (in $\torus^3$) is contained in one of the carriers,
709: and, in the generic case, is everywhere dense in it.
710: Let $D^2_{ls}\subset \real^2_{B,a}$ be planar two-discs
711: orthogonal to the magnetic field such that $\partial
712: D^2_{ls}=\beta_{ls}$. We define the ``closure'' $N_l$ of every carrier
713: $L_l$ as follows: $$N_l=M_l\cup\Big(\bigcup_{s} D^2_{ls}\Big).$$
714: By construction we also have
715: $$N_l\bigcap N_{l'}=\varnothing$$
716: for $l\neq l'$.
717: 
718: We call
719: our system {\it stable topologically completely integrable} if
720: the genus of each surface $N_l $ is equal to one, and this
721: picture is stable under arbitrary small enough perturbations of the magnetic field.
722: 
723: We call the system {\it chaotic} if the genus of some $N_l$ is greater
724: than one. According to the main theorem of \cite{D1}, the latter
725: situation is always topologically unstable.
726: 
727: According to \cite{NM,NM1,MN}, what is important for physical
728: applications, is the following \emph{topological resonance}
729: property of our system. In the stable topologically completely
730: integrable case, all the closures $N_l$ of the carriers of open
731: trajectories have the same up to sign nonzero homology class:
732: $$[N_l]=\pm \mu\in H_2(\torus^3,\integer),\quad\mu\neq 0,$$
733: which is an indivisible element of the group
734: $H_2(\torus^3,\integer)\cong\integer^3$. The number of the tori
735: $N_l\subset \torus^3$ is even because the sum of their homology classes is
736: equal to the class of Fermi surface, which is zero.
737: (Note that every homologically nontrivial connected closed nonselfintersecting
738: two-manifold $M\subset \torus^3$ always represent an indivisible
739: homology class. Any two such submanifolds with empty
740: intersection represent the same homology class up to
741: sign.) The class $\mu\in H_2(\torus^3,\integer)\cong\integer^3$
742: is presented by three relatively prime integers $\mu(B)=(m_1,m_2,m_3)$.
743: 
744: The integral vector $\mu(B)$ remains
745: unchanged under small perturbations of $B$. Therefore, there
746: is an open set on the sphere $S^2$ with the same $\mu(B)$.
747: This set as well as the integral vector $\mu$ is
748: an observable characteristics of our system,
749: and it can be found experimentally by measuring the
750: conductivity tensor in the presence of strong
751: enough magnetic fields having generic directions.
752: The stable topologically integrable
753: case occurs for all directions $B/|B|\in S^2$
754: of the magnetic field from an everywhere dense open subset of
755: $S^2$. It was proved
756: in \cite{D2,D3} (by two different methods)
757: that this picture may be not valid
758: for directions $B/|B|\in S^2$ of the magnetic field from
759: a nonempty subset whose codimension is at least one.
760: 
761: In terms of quasiperiodic functions, we can say that ``non-typical''
762: functions $\varphi_a(p)=\epsilon(p)|_{\real^2_{B,a}}$ with three quasiperiods,
763: \emph{i.e.}, such that open connected
764: components of their level sets don't have a strong asymptotic direction,
765: all lie in a subset that has codimension one (in some
766: natural sense). Examples of level lines with chaotic behavior in
767: the case $n=3$ were constructed in \cite{D2}. We call such level lines
768: \emph{strongly chaotic trajectories}. Interesting attempts were
769: made in order to find physical properties of the conductivity in these
770: cases. For some special examples it was done in work \cite{DM}
771: but in general the stochastic properties of these trajectories are
772: unknown. A.\,Maltsev formulated the following conjecture.
773: 
774: \begin{conjecture}
775: The contribution of strongly chaotic trajectories to the
776: conductivity tensor tends to zero when $|B|$ grows (remaining in
777: a reasonable range), which includes the conductivity in
778: the direction of the  magnetic field itself.
779: \end{conjecture}
780: 
781: Previously, S.\,Tsarev constructed a ``weakly chaotic'' example
782: (unpublished, see work \cite{D2}). In his case, there is a rational dependence
783: between the components of $B$, and there is just one carrier of
784: open trajectories in our sense, which coincides with the Fermi surface.
785: However, the closure of any trajectory in $\torus^3$ is not the whole
786: surface, but just a half of it, which is homeomorphic
787: to a $2$-torus with two holes. The holes are not homologically
788: trivial in $\torus^3$, so they are not regarded as closed in $\real^3$.
789: In Tsarev's example, the nonclosed level lines still have an
790: asymptotic direction in a weaker sense,
791: $$\gamma(t)=(x_0,y_0)+t\cdot(x_1,y_1)+o(t),$$
792: but the projection of any trajectory to a straight line perpendicular to $(x_1,y_1)$
793: is unbounded.
794: 
795: It is interesting to look at the behavior of trajectories in the
796: special (nongeneric) case of the Fermi surface
797: $$\epsilon(p)=\cos(p_1)+\cos(p_2)+\cos(p_3)=0.$$
798: Examples of this type were investigated numerically and
799: analytically in works \cite{D2,RdL}. There are chaotic trajectories for the set of
800: magnetic fields whose Hausdorf dimension is presumably equal to some
801: $\alpha$ with
802: $$1<\alpha<2.$$
803: There are many (in fact, infinitely many)
804: different stable topologically completely
805: integrable zones on the sphere $S^2$ having different integral
806: characteristics $\mu(B)\in\integer^3$. We
807: call this type of examples \emph{generic symmetric
808: levels}, see below.
809: 
810: We conjecture the following.
811: 
812: \begin{conjecture}
813: (i) For a generic connected smooth two-manifold $M_F\subset
814: \torus^3$ homologous to zero, the set of chaotic directions of
815: the magnetic field has Hausdorf dimension less than one in $S^2$.
816: 
817: (ii) For a generic $1$-parametric smooth family $M_{F,t}\subset \torus^3$
818: of such Fermi surfaces this set has Hausdorf
819: dimension less than two.
820: \end{conjecture}
821: 
822: A detailed investigation of this problem containing the proofs of
823: all topological statements needed for physical
824: applications found in \cite{NM,NM1} is performed in
825: \cite{D2,D3}. Special attention is paid there to one-parametric families of
826: Fermi surfaces that are levels of the same Morse function
827: $f:\torus^3\rightarrow \real$:
828: $$M_c=\{f(p)=c\}.$$
829: It is proved that, for any $B$ from a stability zone,
830: open trajectories live on the levels $M_c$ from a connected interval $c_1(B)\leqslant
831: c\leqslant c_2(B)$ of the real line. For a $B$ away from the stability
832: zones, the strongly chaotic
833: behavior might appear only on a single level $c(B)\in\real$.
834: As a corollary we obtain the following result:
835: 
836: \smallskip
837: \centerline{\parbox{0.8\textwidth}{For a function
838: f=$\epsilon(p)$ with symmetry $\epsilon(p+p_0)=-\epsilon(p)$,
839: where $p_0\in\torus^3$ is some shift,
840: strongly chaotic trajectories cannot appear on the levels $M_c$ with
841: $c\neq0$ because otherwise they must appear on $M_{-c}$, too,
842: for the same $B$, which is impossible.
843: In such a case we call the level $c=0$ a \emph{generic symmetric level}.}}
844: \smallskip
845: 
846: According to
847: our conjecture, the Hausdorf dimension of the set of $B/|B|\in S^2$ for which
848: the strongly chaotic behavior occurs on such a level is equal to some $\alpha<2$.
849: 
850: The surface $\sum_{j=1}^{j=3} \cos(p_j)=0$ gives an example of a generic
851: symmetric level with $p_0=(\pi,\pi,\pi)$.
852: 
853: Some more details about chaotic
854: trajectories and stability zones for the case of three
855: quasiperiods will be given below. They will be needed for
856: proving our main
857: result about quasiperiodic functions with four quasiperiods (see the next
858: section).
859: 
860: \section{The stable topological complete integrability for $n=4$
861: quasiperiods}
862: 
863: Let us consider now the case of $n=4$ (or more) quasiperiods.
864: For every direction $\Pi$ of two-planes in $\real^n$ (\emph{i.e.}, a two-dimensional
865: vector subspace $\Pi\subset\real^n$), the
866: original $n$-periodic Morse function
867: $f:\torus^n\rightarrow \real$
868: defines a family of descendants $\{\varphi_a(y)\}$ on the family
869: affine two-planes $\real^2_{\Pi,a}\subset
870: \real^n$ having direction $\Pi$.
871: 
872: We call the level $\{f=c\}$
873: of the function $f$ \emph{topologically completely
874: integrable} (\emph{TCI}) for the direction $\Pi$ if, for each
875: $\varphi_a$ from the family, all regular connected components of the level
876: line $\varphi_a(y)=c$ are either compact or have a strong asymptotic
877: direction. We call this level \emph{stable
878: TCI} if this property remains unchanged under small
879: perturbations of the function $f$ and the direction $\Pi$,
880: which is a point in the Grassmanian manifold $G_{n,2}$.
881: 
882: We say that
883: the Stable TCI level satisfies the {\it topological
884: resonance} condition (for given $\Pi$) if there exists an integral hyperplane
885: $\mu\subset \real^n$, $\mu\cap\integer^n\cong\integer^{n-1}$,
886: such that all open regular trajectories have
887: the same asymptotical direction $\eta_\Pi$ that coincides with
888: the direction of the straight line $\mu\cap\Pi\cong\real$.
889: Since $\mu$ is integral, it must remain unchanged under small perturbations of
890: anything.
891: 
892: Let us point out that even the ``trivial case'' $n=2$ is
893: meaningful (as a subject of the elementary Morse theory on the
894: $2$-torus): for a generic double-periodic function on the plane
895: there exists a level $f=c$ with a connected component presenting a nontrivial
896: indivisible
897: homology class $\mu\in H_1(\torus^2,\integer )$.
898: All other components of every level are
899: either homologically trivial or homologous to $\pm\mu$. For
900: Morse functions with exactly four critical points and critical values
901: $c_0<c_1<c_2<c_3$, all levels $f=c$, with $c_1<c<c_2$, have exactly two
902: connected components, whose homology classes are $\pm\mu$. All other nonsingular
903: levels are either compact or empty.
904: 
905: \begin{description}
906: \item[Question]
907: Consider famous real nonsingular quasiperiodic
908: finite-gap solutions of the KdV equation
909: $$u(x,t)=2\partial_x^2\log\Theta (xU+tV+U_0) + c_{\Gamma}$$ with
910: an arbitrary number of quasiperiods (or gaps). Are the levels
911: $u(x,t)=c$ always stable topologically completely integrable or
912: they can be chaotic? How to find their strong asymptotic
913: direction and their integer-valued characteristic $\mu$?
914: \end{description}
915: 
916: P.\,Grinevich pointed out to us that, for real smooth finite-gap
917: solutions of the KdV equation,
918: the $n$-periodic function $f=2\partial_U^2 \log
919: \Theta(\eta_1,\dots,\eta_n)+c_{\Gamma}$ on the $\eta$-space
920: is always a Morse
921: function on the real $n$-torus with $2^n$ critical points, simply
922: because it can be reduced to the form
923: $f=\sum_{j=1}^{j=n}\alpha_j\sin x_j$ by a diffeomorphism of the
924: torus isotopic to the identity. There is a canonical  lattice in the
925: $\eta$-space generated by the so-called $a$-cycles which are the
926: real finite gaps of the 1D Schr\"odinger operator on a hyperelliptic
927: ``spectral'' Riemann surface $\Gamma$. The real constants
928: $\alpha_j$ depend on the spectrum. As a conclusion, we get the
929: following:
930: 
931: \smallskip
932: \centerline{\parbox{0.8\textwidth}{For
933: generic real nonsingular two-gap solutions of the KdV equation, there
934: exist a critical value $c_{\mathrm{cr}}$ such that all constant speed
935: levels $u(x,t)=c$,
936: $$c_1=c_{\Gamma}-c_{\mathrm{cr}}<c<c_{\Gamma}+c_{\mathrm{cr}}=c_2$$
937: are periodic
938: perturbations  of a family of straight lines with integral
939: direction $m_1:m_2$ on the plane
940: with lattice. This direction is locally rigid, but globally depends
941: on the constants $\alpha_j$. All other levels are either compact
942: or empty. We call it the \emph{topological speed} of the solution.}}
943: \smallskip
944: 
945: The computational studies of this problem for finite-gap solutions
946: are now being investigated numerically.
947: 
948: We concentrate now on the case $n=4$ quasiperiods. Work
949: \cite{N2} presents an idea of the proof that, for every generic Morse
950: function $f$ and a noncritical generic level $f=c$, there exists an open
951: everywhere dense set of two-plane directions
952: $\Pi\in G_{4,2}$ for which the level $f=c$
953: is stable TCI. The proof of this theorem
954: requires the use of some extension of the results of work
955: \cite{D3}. Here we make some corrections to the statement
956: of~\cite{N2} and provide a complete proof.
957: 
958: We believe that a generic level is stable TCI for all directions $\Pi$
959: from a subset $S\subset G_{4,2}$ whose measure is full in $G_{4,2}$. However,
960: we don't have an idea how to prove this conjecture. For $n>4$
961: nothing like that is expected.
962: 
963: Now we start a detailed investigation of the case $n=4$. Even
964: Question~1 of the previous section presents a difficulty here.
965: It is possible that a
966: single level set of a quasiperiodic function with four
967: quasiperiods contains a family of compact components without an
968: upper bound for their size. An example can be constructed easily.
969: 
970: However, we are going to show that there is an open everywhere
971: dense open set of
972: quasiperiodic functions $\varphi$ with four quasiperiods such that
973: the level lines $\varphi=\const$ have the same qualitative
974: behavior as those in the typical case of three quasiperiods.
975: The precise formulation is as follows.
976: 
977: \begin{theorem}\label{th1}
978: There exists an open everywhere dense
979: subset $S\subset C^\infty(\torus^4)$ of $4$-periodic
980: functions $f$ and an open everywhere dense subset
981: $X_f\subset G_{4,2}$ depending on $f$ such that
982: any level $M^3_c=\{f=c\}$ of $f$ is stable TCI (or does not
983: contain open trajectories at all) for any $\Pi\in X_f$.
984: 
985: Moreover, for any regular open trajectory, the remainder
986: term $O(1)$ in (\ref{strong}) as well as
987: the diameter of any compact trajectory are bounded from above by a constant $C$
988: not depending on the affine plane $\real^2_{\Pi,a}$ containing the trajectory,
989: provided that the level $c$ and the direction $\Pi\in X_f$ are fixed.
990: \end{theorem}
991: 
992: Let us make a remark about notation and terminology.
993: Once we switched to the case of four quasiperiods,
994: our problem is no longer relevant to the discussed above physical model
995: of conductivity in normal metals in the presence of a magnetic field.
996: So we change the notation for the coordinates in $\real^n$ from $p_l$, which
997: we used for quasimomentum, to more customary, $x_l$, $l=0,1,2,3$,
998: and don't longer think
999: of the lattice $\integer^4\subset\real^4$ as the one dual to
1000: some physical lattice. We think of the ``magnetic field''
1001: $B$ as a linear mapping from $\real^4$ to $\real^2$ (or from $\real^3$
1002: to $\real$ in the $n=3$ case) such that $\Pi=\ker(B)$.
1003: Thus, by $\real^2_{\Pi,a}$ we mean the two-plane $B^{-1}(a)$,
1004: where $a\in\real^2$. In the case $n=3$ we may also think of $B$ as a vector
1005: perpendicular to the plane $\Pi$.
1006: However, we keep calling connected components of
1007: the intersections of $M_c^3$ with the two-planes $\real^2_{\Pi,a}$
1008: trajectories, just for briefness.
1009: 
1010: We start by recalling results of~\cite{D2,D3} for the
1011: three-dimensional case in the form needed to prove our theorem. Let
1012: $B:\real^3\rightarrow\real$ be a linear function of irrationality
1013: degree three, \emph{i.e.}, of the form $B(x)=B_1x_1+B_2x_2+B_3x_3$,
1014: where $B_1$, $B_2$, $B_3$ are reals linearly independent over $\integer$, and
1015: let $f:\torus^3\rightarrow\real$ be a generic smooth function. By
1016: `generic' we mean that $f$ does not satisfy certain conditions
1017: that have codimension $\geqslant1$. However, we shall pay
1018: attention to codimension one singularities as, in order to deal
1019: with the four-dimensional case, we are going to consider
1020: one-parametric families of three-dimensional pictures.
1021: 
1022: We abuse notation by using the same letter $f$ for the lift of $f$
1023: to the covering $3$-space $\real^3$. We use notation $M^2_c$
1024: for the level set $f^{-1}(c)$ in
1025: $\torus^3$ and $\widehat{M^2_c}$ for its cover in $\real^3$. By
1026: $\gamma_{a,c}$ we denote the whole intersection of $\widehat{M^2_c}$
1027: with the plane $\real^2_{\Pi,a}=B^{-1}(a)$. So, the trajectories that we are studying are
1028: regular connected components of $\gamma_{a,c}$ or their projections to $\torus^3$.
1029: 
1030: First of all, consider closed trajectories on $M^2_c$. Notice
1031: that, since we assumed $B$ to be of maximal irrationality degree,
1032: a trajectory in $\real^3$ is closed if and only if so is its image
1033: in $\torus^3$. Without the assumption on $B$ this may be not true,
1034: since a closed trajectory in $\torus^3$ may then be
1035: non-homologous to zero, in which case its cover in $\real^3$
1036: consists of infinite ``periodic'' trajectories treated as ``open''
1037: in the physical applications.
1038: 
1039: For a generic $f$, compact trajectories on every $M^2_c$ form
1040: finitely many cylinders whose bases are either saddle connections
1041: or extrema of the restriction $B|_{M^2_c}$. Obviously, the length
1042: of compact trajectories is bounded from above by some constant.
1043: 
1044: Let $U$ be the set of $c$ such that $\gamma_{a,c}$ has unbounded
1045: connected components for some $a$. In other words, $c\in U$ if and
1046: only if $M^2_c$ contains open trajectories that are not saddle
1047: connections. The following picture, which was sketched in
1048: the previous section, can be extracted from work
1049: \cite{D3}:
1050: 
1051: \smallskip
1052: \noindent\hskip 0.1\textwidth{\parbox{0.8\textwidth}{The
1053: set $U$ is either a closed interval, $U=[c_-,c_+]$, or
1054: just one point, $U=\{c_0\}$.
1055: 
1056: If $U=[c_-,c_+]$ is a nontrivial interval, then for any $c\in U$,
1057: there is a (unique) family of two-tori $\torus^2_{c,1},\ldots,
1058: \torus^2_{c,2k}$ (with $k$ depending on $c$)
1059: imbedded in $\torus^3$ such that
1060: \begin{enumerate}
1061: \item Each $\torus^2_{c,i}$ consists of the closure of some
1062: open trajectory on $M^2_c$ and a few (may be zero) planar
1063: disks perpendicular to $B$;
1064: \item Every open trajectory is contained by whole in one of
1065: the tori $\torus^2_{c,i}$;
1066: \item All the tori $\torus^2_{c,i}$ define the same up to sign
1067: nonzero homology class $\mu$ in $H_2(\torus^3,\integer)$;
1068: \item For all but finitely many $c$, the tori $\torus^2_{c,i}$
1069: are pairwise disjoint, and, in this case, a sufficiently
1070: small variation of $c$ causes small deformation of the tori.
1071: For the exceptional $c$'s they can be
1072: made disjoint by a small perturbation. At such a $c$, a couple of tori
1073: is born or killed.
1074: \end{enumerate}
1075: All the picture is stable in this case, which means that after a
1076: small enough perturbation of $f$, the interval $U$ and the family
1077: of tori $\torus^2_{c,i}$ are perturbed slightly. In particular,
1078: the homology class $\mu$ fixed.}}\hfill\\
1079: 
1080: \begin{remark}
1081: In the setting of papers~\cite{D2,D3},
1082: the function $f$ was assumed to be fixed
1083: and the point of concern was the dependence of the behavior of
1084: our dynamical system on the magnetic field $B$ and on the
1085: level of the function $f$. The stability
1086: of the whole picture under small perturbations of the function $f$
1087: was not discussed. However, the arguments of those works can be
1088: easily modified in order to prove such stability. Indeed,
1089: one of the key observations in~\cite{D2,D3} is that, locally,
1090: the qualitative behavior of the trajectories (including the existence
1091: of a strong asymptotic direction) depends only
1092: on finitely many parameters, which are certain critical values of
1093: the ``height'' function $B(x)$ restricted
1094: to the surface $\{f=\const\}$. (For instance,
1095: the existence of strongly chaotic examples
1096: was proved in~\cite{D2} by specifying the combinatorial
1097: structure of the surface and particular values of the
1098: parameters.) It is easy to see that those parameters
1099: behave nicely under small perturbations of $f$,
1100: so extending the arguments of~\cite{D2,D3}
1101: to this, more general type of perturbations requires
1102: almost no additional work.
1103: \end{remark}
1104: 
1105: Let us describe the three-dimensional picture in more details. For
1106: a generic level surface $M^2\subset\torus^3$, the structure of
1107: trajectories on $M^2$ is as follows. Compact trajectories form a
1108: few open cylinders whose ``ends'' approach either an extremum
1109: point of the function $f|_{M^2}$ or a saddle connection cycle, see
1110: Fig.~\ref{cylinders}.
1111: \begin{figure}[ht]
1112: \centerline{\epsfig{file=cc1.eps,height=100pt}\hskip1cm
1113: \epsfig{file=cc2.eps,height=100pt}\hskip1cm\epsfig{file=cc3.eps,height=100pt}}
1114: \caption{Cylinders of compact
1115: trajectories}\label{cylinders}
1116: \end{figure}
1117: The rest of the surface (if not empty) consists of an
1118: even number of two-tori with or without holes, and each hole is a
1119: saddle connection cycle. Each hole can be glued up by a planar
1120: disk perpendicular to the vector $B$. We denote the obtained
1121: surface by $N$. The preimage $\widehat N\subset\real^3$ of $N$
1122: under the projection $\nu:\real^3\rightarrow\torus^3$ is a family of
1123: finitely deformed periodic ``wrapped''  planes in $\real^3$, see
1124: Fig.~\ref{wrapped}.
1125: \begin{figure}[ht]
1126: \centerline{\epsfig{file=wp.eps,height=100pt}}
1127: \caption{A wrapped plane}\label{wrapped}
1128: \end{figure}
1129: 
1130: What happens to $\widehat N$ when the surface $M^2$ changes?
1131: Small deformations of the surface $M^2$ cause just small
1132: deformations of the tori and their covering planes. Suppose we
1133: have a generic $1$-parametric family of surfaces $M^2(t)$. This
1134: means that we consider a generic $1$-parametric family of
1135: functions $f_t:\torus^3\rightarrow\real$, and for each $t$, the
1136: surface $M^2(t)$ is defined by the equation $f_t(x)=\const$.
1137: 
1138: When the parameter $t$ varies, the connected components of $N$ are
1139: just deformed while they stay apart from each other. However,
1140: eventually two tori can collide and disappear or, on the contrary,
1141: a pair of tori can be born. This occurs when $M^2(t)$ traverses a
1142: subset which has codimension one in a natural sense. It is not
1143: important here whether $M^2(t)$ is the family of level surface of
1144: a single function or an arbitrary generic one-parametric family of
1145: surfaces.
1146: 
1147: The generic tori collision was described in~\cite{D3}.
1148: It was assumed in \cite{D3} that the family of
1149: surfaces $M^2(t)$ is the family of level surfaces of the same
1150: function, $M^2(t)= \{x\in\torus^3\;|\;f(x)=t\}$. However, the
1151: argument is exactly the same for an arbitrary generic family of
1152: surfaces.
1153: 
1154: The following two types of tori collision are possible
1155: in the generic case.
1156: \begin{enumerate}
1157: \item
1158: A cylinder of closed trajectories with bases attached to two
1159: different components of $N$ collapses. The corresponding
1160: codimension-one condition has the following form: two different
1161: saddles get joined by a saddle connection.
1162: This causes an ``interaction'' of pairs of open
1163: trajectories lying on the collided tori, which turns
1164: them into infinitely many closed trajectories,
1165: see Fig.~\ref{degenerate-cylinder}.
1166: \begin{figure}[ht]
1167: \centerline{\epsfig{file=ccc1.eps,height=200pt}
1168: \hskip0.5cm\raisebox{100pt}{$\rightarrow$}\hskip0.5cm
1169: \epsfig{file=ccc2.eps,height=200pt}
1170: \hskip0.5cm\raisebox{100pt}{$\rightarrow$}\hskip0.5cm
1171: \epsfig{file=ccc3.eps,height=200pt}}
1172: \caption{Collapse of a cylinder}
1173: \label{degenerate-cylinder}
1174: \end{figure}
1175: \item
1176: A Morse-type surgery occurs on $M^2$ that results in a one-handle added
1177: to the surface. The behavior of trajectories is shown in Fig.~\ref{morse}.
1178: \begin{figure}[ht]
1179: \centerline{\epsfig{file=ms1.eps,height=200pt}
1180: \hskip0.5cm\raisebox{100pt}{$\rightarrow$}\hskip0.5cm
1181: \epsfig{file=ms2.eps,height=200pt}
1182: \hskip0.5cm\raisebox{100pt}{$\rightarrow$}\hskip0.5cm
1183: \epsfig{file=ms3.eps,height=200pt}}
1184: \caption{A Morse surgery destroying open trajectories}\label{morse}
1185: \end{figure}
1186: \end{enumerate}
1187: 
1188: It is important to note here that whenever $N$ consists of just
1189: two tori and $M^2(t)$ passes a singularity of one of the two types
1190: mentioned above, then all open trajectories get destroyed, so that
1191: $N$ is empty right after the critical event. However, the
1192: following is not proven to be impossible in a generic
1193: one-parametric family of surfaces: bearing and canceling of a
1194: pair of tori occurs alternatingly at moments $t_1,t_2,t_3,\dots$
1195: so that the sequence $(t_n)$ converges to some $t_*$ and an
1196: ergodic regime occurs on $M^2(t_*)$. The latter means that there
1197: is an open trajectory on $M^2(t_*)$ whose closure has genus more
1198: than one (actually, it should then be equal to three). The
1199: existence of such ergodic regimes was proven in~\cite{D2,D3}, and it
1200: was shown only that such regimes satisfy a codimension one
1201: condition. However, we will not need to deal with ergodic regimes
1202: in order to prove our result.
1203: 
1204: Now we turn to the case when $M^2(t)=M^2_t$ is the family of level
1205: surfaces of a generic function on $\torus^3$.
1206: 
1207: Consider the restriction of the function $f$ to the plane $\real^2_{\Pi,a}$
1208: for some $a$. Let $V\subset\real^2_{\Pi,a}$ be the union of all compact
1209: components of $\gamma_{a,c}$ over all $c$, and $W\subset\real^2_{\Pi,a}$ the
1210: union of all unbounded components of $\gamma_{a,c}$. We have
1211: $\real^2_{\Pi,a}=V\cup W$, $V\cap W=\varnothing$, $V$ is open. Notice:
1212: connected components of $V$ are not necessarily bounded. Let
1213: $V_1,V_2,\ldots$ be the connected components of $V$. It is easy to
1214: see that $f$ is constant on $\partial V_i$ for any $i$.
1215: 
1216: For $x\in\real^2_{\Pi,a}$ we put
1217: $$\overline f(x)=\left\{\begin{array}{ll}f(x)&\text{if }x\in W,\\
1218: f(\partial V_i)&\text{if }x\in V_i.\end{array}\right.$$
1219: By doing so for all $a$, we obtain a new function $\overline f:
1220: \real^3\rightarrow\real$. We use the following notation:
1221: $$N_c=\{x\in\torus^3\;;\;\overline f(x)=c\}.$$
1222: The function $f$ and its level sets $N_c$
1223: have the following properties.
1224: 
1225: \begin{lemma}
1226: The function $\overline f$ is a well defined continuous function
1227: on $\torus^3=\real^3/\integer^3$.
1228: 
1229: If $U=\{c_0\}$, then $\overline f\equiv c_0$.
1230: 
1231: If $U=[c_-,c_+]$, where $c_-<c_+$, then for all but finitely many
1232: $c\in(c_-,c_+)$, we have
1233: $$N_c=\bigcup\limits_i\torus^2_{c,i}.$$
1234: 
1235: For all $c\in[c_-,c_+]$ a small regular neighborhood of
1236: $N_c$ is homeomorphic to the union of a few copies of
1237: $\torus^2\times[0,1]$.
1238: \end{lemma}
1239: 
1240: \begin{proof}
1241: In the case $U=\{c_0\}$ our claim is trivial.
1242: 
1243: Assume that $U=[c_-,c_+]$ with $c_-<c_+$.
1244: By construction, $\torus^2_{c,i}\cap M_c$ consists of open trajectories,
1245: thus, we have $\overline f(x)\equiv c$ on $\torus^2_{c,i}\cap M_c$.
1246: The whole torus $\torus^2_{c,i}$ is obtained from
1247: $\torus^2_{c,i}\cap M_c$ by attaching disks each of which
1248: lies in the plane $\real^2_{\Pi,a}$ for some $a$. The boundary of such a disk is
1249: a part of a singular unbounded component of the level set
1250: of $f|_{\real^2_{\Pi,a}}$. By construction, we have $\overline f\equiv c$
1251: in such a disk. Therefore, we always have
1252: $\torus^2_{c,i}\subset N_c$.
1253: 
1254: Let us look at what happens with the tori $\torus^2_{c,i}$ when
1255: $c$ varies. For all but finitely many $c$ the tori
1256: $\torus^2_{c,i},\torus^2_{c,j}$ are disjoint if $i\ne j$.
1257: Moreover, for any $c\ne c'$ and any $i,j$, the tori $\torus^2_{c,i}$
1258: $\torus^2_{c',j}$ are always disjoint.
1259: 
1260: When $c$ varies, tori $\torus^2_{c,i}$ are continuously deformed
1261: except at a few values of $c$, where one of the following happens:
1262: 1) two tori collide and then disappear; 2) two tori are
1263: newly born. The latter event is opposite to the first one.
1264: 
1265: Let us describe torus collision in more detail.
1266: At the moment of the collision we have a closed domain $W$
1267: in $\torus^3$ that has the form of the manifold
1268: $\torus^2\times[0,1]$ in which some intervals
1269: $x\times[0,1]$ are collapsed to a point. There may be
1270: just one such points $x$ or a closed disk $D^2\subset\torus^2$
1271: of such points. The first case corresponds to an index one or index two
1272: Morse critical point of $f$, whereas the latter
1273: corresponds to a degenerate cylinder of closed trajectories.
1274: 
1275: The interior of the domain $W$ is filled by compact trajectories
1276: and, by construction, the function $\overline f$ is constant
1277: inside $W$. Thus, $W$ is a connected component of some $N_c$,
1278: since $W$ is squeezed between the two collided tori. We call such
1279: a $W$ \emph{pseudotorus}.
1280: 
1281: So, we have the following picture. The decomposition of $\torus^3$
1282: into the union of (connected components of) $N_c$ over all $c$
1283: is nothing else but a trivial fibration over $S^1$ with fibre $\torus^2$,
1284: with a few fibres replaced by pseudotori.
1285: 
1286: Schematically this is shown in Fig.~\ref{pseudotori}.
1287: \begin{figure}[ht]
1288: \centerline{\epsfig{file=pseudot.eps,height=150pt}}
1289: \caption{A family of tori with a few replaced by pseudotori}
1290: \label{pseudotori}
1291: \end{figure}
1292: 
1293: \end{proof}
1294: 
1295: Now we return to the four-dimensional case. Let $\Pi\in G_{4,2}$
1296: be a two-plane defined by a linear mapping $B:\real^4\rightarrow\real^2$.
1297: By $\real^2_{\Pi,a,b}$ we denote the affine plane $B^{-1}(a,b)\subset\real^4$,
1298: and by $M^4_{\leqslant c}$ (respectively, $M^4_{\geqslant c}$)
1299: the subset of $\torus^4$ defined by the inequality $f(x)\leqslant c$
1300: (respectively, $f(x)\geqslant c$).
1301: 
1302: Let $N\subset\torus^4$ be a submanifold (or, more generally, a subset).
1303: We say that $N$ is \emph{essentially below} (respectively, \emph{essentially
1304: above}) $M^3_c$ if for any $a,b\in\real$,
1305: the intersection $\widehat N\cap\real^2_{\Pi,a,b}$
1306: is disjoint from all unbounded components of $\widehat M^4_{\geqslant c}
1307: \cap\real^2_{\Pi,a,b}$ (respectively, of $\widehat M^4_{\leqslant c}\cap\real^2_{\Pi,a,b}$).
1308: Thus, the property of $N$ to be essentially below $M^3_c$ depends on $\Pi$.
1309: 
1310: The following two facts are proved by analogy with the 3D case.
1311: 
1312: \begin{lemma}
1313: If $N$ is essentially below or essentially
1314: above $M^3_c$ for a given $\Pi$ then this remains
1315: true after a small perturbation of $\Pi$, $f$, and $c$.
1316: \end{lemma}
1317: 
1318: \begin{lemma}
1319: If there exists a homologically nontrivial $3$-torus $N$
1320: which is essentially above or essentially below $M^3_c$,
1321: then the assertion of Theorem~\ref{th1} is true
1322: for these specific $f$, $\Pi$, and $c$.
1323: \end{lemma}
1324: 
1325: Thus, in order to prove Theorem~\ref{th1} it suffices to
1326: show that for everywhere dense set of pairs $(f,\Pi)$,
1327: and for each $c$,
1328: there exists a homologically nontrivial $3$-torus $N\subset\torus^4$
1329: which is essentially below or essentially above $M^3_c$.
1330: 
1331: Let $B=(\ell_1,\ell_2)$ be a couple of
1332: linear functions on $\real^4$ such that
1333: \begin{enumerate}
1334: \item
1335: the function $\ell_1$ is rational, i.e., $\ell_1\in(\integer^4)^*$;
1336: \item
1337: the restriction of $\ell_2$ to the integral three-plane $\ell_1=0$
1338: has irrationality degree three.
1339: \end{enumerate}
1340: Obviously, the set of $2$-planes $\Pi=\ker B$ defined by
1341: $\ell_1,\ell_2$ of this form is everywhere dense in $G_{4,2}$.
1342: 
1343: Without loss of generality, we assume that $\ell_1(x)=x_0$,
1344: $\ell_2(x)=H_1x_1+H_2x_2+H_3x_3$, where
1345: $x=(x_0,x_1,x_2,x_3)\in\real^4$, $\qopname\relax
1346: o{rank}_\integer\langle H_1,H_2,H_3\rangle=3$. We consider the
1347: $4$-torus $\torus^4$ as a one-parametric family of three-tori
1348: $\torus^3_t=\{x_0=t\}$. For any $t\in[0,1]$, we deal with the
1349: restrictions $f_t$ and $\ell_{2,t}$ of respectively $f$ and
1350: $\ell_2$ to $\torus^3_t$ as in the three-dimensional case. We
1351: introduce $U_t=[c_{t-},c_{t+}]$, $\overline{f_t}$, $N_{t,c}$ as
1352: before.
1353: 
1354: Let us consider the dependence of the interval $U_t$ on $t$.
1355: The endpoints $c_{t\pm}$ of the interval $U_t$ are continuous functions
1356: of $t$. Moreover, in the regions where $c_{t+}>c_{t-}$
1357: these functions are piecewise smooth. This follows from
1358: the fact that locally, near a generic $t$, they are defined
1359: by a condition of the form: two saddles on $M^2_{t,c_{\pm}}$
1360: are connected by a separatrix. Here we call such intervals \emph{stability zones}.
1361: To every stability zone there corresponds an integral vector
1362: $\mu\in H_1(\torus^3,\integer)=\integer^3$, which we call
1363: the \emph{label} of the zone.
1364: 
1365: Figure~\ref{cpm} shows how the functions $c_{t\pm}$ may look like
1366: in the generic case.
1367: It is possible that at some $t$ we have $c_{t+}=c_{t-}$,
1368: see Fig.~\ref{cpm}a). This may
1369: occur at the boundary of a stability zone or at $t$ such that
1370: the open trajectories in $M^3_t$ are chaotic.
1371: \begin{figure}[p]
1372: \noindent a)
1373: 
1374: \centerline{\begin{picture}(425,160)
1375: \put(0,30){\epsfig{file=cpma.eps,width=400pt}}
1376: \put(0,10){\vector(1,0){425}} \put(422,0){$t$}
1377: \put(150,80){$c_{t-}$} \put(150,130){$c_{t+}$}
1378: \put(402,71){$c_0(t)$}
1379: \end{picture}}
1380: 
1381: 
1382: \vskip1cm
1383: \noindent b)
1384: 
1385: \centerline{\begin{picture}(425,160)
1386: \put(0,30){\epsfig{file=cpmb.eps,width=400pt}}
1387: \put(0,10){\vector(1,0){425}} \put(422,0){$t$}
1388: \put(150,76){$c_{t-}$} \put(150,135){$c_{t+}$}
1389: \multiput(0,152)(10,0){41}{\line(1,0)5}
1390: \put(408,149){$(c_-)_{\mathrm{max}}$}
1391: \multiput(0,99)(10,0){41}{\line(1,0)5}
1392: \put(408,96){$(c_+)_{\mathrm{min}}$}
1393: \end{picture}}
1394: 
1395: \vskip1cm
1396: \noindent c)
1397: 
1398: \centerline{\begin{picture}(425,160)
1399: \put(0,30){\epsfig{file=cpmc.eps,width=400pt}}
1400: \put(0,10){\vector(1,0){425}} \put(422,0){$t$}
1401: \put(150,62){$c_{t-}$} \put(150,145){$c_{t+}$}
1402: \multiput(0,106)(10,0){41}{\line(1,0)5}
1403: \put(408,89){$(c_-)_{\mathrm{max}}$}
1404: \multiput(0,92)(10,0){41}{\line(1,0)5}
1405: \put(408,103){$(c_+)_{\mathrm{min}}$}
1406: \end{picture}}
1407: \caption{Functions $c_{t\pm}$ in the generic case}\label{cpm}
1408: \end{figure}
1409: For all such $t$ we have $c_{t+}=c_{t-}=c_0(t)$, where $c_0(t)$ is
1410: a piecewise smooth function of $t$. It is defined locally by a
1411: condition of the form: the sum of ``heights'' of certain saddles
1412: equals to zero, see~\cite{D2,D3}.
1413: 
1414: It is most likely that any ``chaotic" $t$ must be an accumulating
1415: point of an infinite sequence of stability zones. In other words,
1416: it cannot happen that the equality $c_{t+}=c_{t-}$ holds
1417: everywhere in a nontrivial interval $(t_1,t_2)$. However, this
1418: does not follow directly from the previous works~\cite{D2,D3}, and
1419: will not be used here.
1420: 
1421: \begin{lemma}\label{=/=}
1422: The equality
1423: \begin{equation}\label{c+=c-}
1424: \min_tc_{t+}=\max_tc_{t-}
1425: \end{equation}
1426: does not hold for a generic $f$.
1427: \end{lemma}
1428: 
1429: \begin{proof}
1430: We should consider the following three cases.
1431: 
1432: Case 1. For all $t$ we have $c_{t+}>c_{t-}$. Then there exists
1433: a smooth periodic function $g(t)$ such that
1434: $c_{t+}>g(t)>c_{t-}$. Condition~(\ref{c+=c-})
1435: will not hold if we disturb the function $f$ in the following way:
1436: $f(t,x_1,x_2,x_3)\mapsto f_\varepsilon(t,x_1,x_2,x_3)=f(t,x_1,x_2,x_3)+
1437: \varepsilon g(t)$, where $|\varepsilon|$ is sufficiently small.
1438: So,~(\ref{c+=c-}) impose a codimension one condition on $f$ in this case.
1439: 
1440: Case 2. For two different $t=t_1,t_2$ we have $c_{t+}=c_{t-}$.
1441: Then condition~(\ref{c+=c-}) will not hold after an arbitrary
1442: perturbation $f\mapsto f+\varepsilon g(t)$, where $g(t)$
1443: is an arbitrary function with $g(t_1)\ne g(t_2)$.
1444: 
1445: Case 3. There is exactly one $t=t_0$ such that $c_{t+}=c_{t-}=c_0$ holds,
1446: and we have $c_{t+}>c>c_{t-}$ for all $t\ne t_0$.
1447: Let us take $t$ close to $t_0$. The interval $[c_{t-},c_{t+}]$
1448: is then small, which means that, when $c$ varies from
1449: $(c_{t-}-\delta)$ to $(c_{t+}+\delta)$ with $\delta>0$,
1450: a pair of tories $N_{t,c}$ is born at $c=c_{t-}$ and then almost
1451: immediately destroyed at $c=c_{t+}$. Therefore,
1452: there are two cylinders of closed trajectories on $M_{t,c_0}$
1453: of very small height.
1454: 
1455: Now let us fix $c=c_0$ and vary $t$. When $t$ approaches $t_0$
1456: from the left, say, we will have two closed trajectory cylinders
1457: that get degenerate at the moment $t=t_0$. When $t$ passes $t_0$,
1458: two cylinders must appear again. The main point here is that
1459: those, new, cylinders appear from \emph{the same} pair of degenerate
1460: cylinders. Indeed, the pair of degenerate cylinders that we obtain
1461: when $t$ approaches $t_0$ from the left cuts $M_{c_0,t_0}$
1462: into two tori. Since the irrationality degree of $\ell_2$
1463: is equal to three, one can show that there may no other degenerate cylinder on
1464: $M_{t_0,c_0}$.
1465: 
1466: Thus, we have the following picture. When $t$ goes from $t_0-\delta$
1467: to $t_0+\delta$, two closed trajectory cylinders degenerate
1468: and then regenerate again. Let $h_1(t)$, $h_2(t)$ be there heights.
1469: So, we not only have $h_1(t_0)=h_2(t_0)=0$, but also
1470: $h_1'(t_0)=h_2'(t_0)=0$, which impose a codimension two
1471: condition on the function $f$.
1472: \end{proof}
1473: 
1474: By construction we have
1475: 
1476: \begin{lemma}\label{<>}
1477: For any $t$ and $c>c_{t+}$ (respectively, $c<c_{t-}$),
1478: the torus $\torus^3_t$ is essentially
1479: below (respectively, essentially above) $M^3_c$.
1480: \end{lemma}
1481: 
1482: 
1483: According to Lemma~\ref{=/=} only the following two cases are possible:
1484: 1) $\min_tc_{t+}<\max_tc_{t-}$; 2) $\min_tc_{t+}>\max_tc_{t-}$.
1485: In Case~1, for any $c$ we have either $c>\min_tc_{t+}$
1486: or $c<\max_tc_{t-}$, and by Lemma~\ref{<>} we are done.
1487: 
1488: So, it remains to consider Case~2,
1489: $\min_tc_{t+}>\max_tc_{t_-}$. This inequality means,
1490: in particular, that the intervals $U_t$
1491: have a nontrivial intersection $U=\cap_tU_t=[c_-,c_+]$, and we
1492: have just one stability zone, which covers the whole circle $S^1$.
1493: This is illustrated in Fig.~\ref{cpm}c).
1494: 
1495: We define a function $\overline f:\torus^4\rightarrow\real$
1496: as in the three-dimensional case by considering
1497: intersections $M_c^3\cap\real^2_{\Pi,a,b}$, and introduce notation
1498: $$N_c=\{x\in\torus^4\;;\;\overline f(x)=c\}.$$
1499: By construction, $\overline{f_t}$
1500: coincides with the restriction of $\overline f$ to $\torus^3_t$, and we
1501: have
1502: $$N_c=\bigcup\limits_tN_{t,c}.$$
1503: 
1504: 
1505: For almost
1506: any $c$, $t$, the intersection $N_c$ with $\torus^3_t$ is either empty
1507: or consists
1508: of $2$-tori, and all those tori have the same up to sign homology class
1509: $\alpha\in H_2(\torus^3,\integer)$.
1510: In this case, the whole torus $\torus^4$ has the structure
1511: of a trivial $\torus^2$-bundle over $\torus^2$ with a $1$-parametric
1512: family of fibres replaced by pseudotori.
1513: \begin{figure}[ht]
1514: \centerline{\begin{picture}(400,320)
1515: \put(0,20){\epsfig{file=t2.eps,height=300pt}}
1516: \put(0,10){\vector(1,0){310}} \put(305,0){$t$} \put(400,160){\hbox
1517: to0pt{\hss Pseudotori}} \put(340,170){\vector(-1,1){55}}
1518: \put(340,160){\vector(-2,-3){37}}
1519: \end{picture}}
1520: \caption{Level lines of $\overline f$ on $\torus^2$}\label{t2}
1521: \end{figure}
1522: The function
1523: $\overline f$ is constant over each fibre, so it can be considered
1524: as a function on $\torus^2$.
1525: 
1526: Figure~\ref{t2} illustrates the structure of level lines of $\overline f$
1527: viewed as a function on $\torus^2$. The preimage of a generic point
1528: is a $2$-torus imbedded in $\torus^4$, and all those $2$-tori are
1529: ``parallel". Whenever the function $\overline f$ has an extremum on
1530: the line $t=\mathrm{const}$, the preimage of the critical point
1531: is a pseudotorus.
1532: 
1533: Including a pseudotorus into a $1$-parametric family of $2$-tori
1534: does not change the topological type of the union of the tori. Indeed,
1535: as we mentioned above, a small regular neighborhood of a pseudotorus
1536: in $\torus^3$ is homeomorphic to $\torus^2\times(0,1)$. In other words,
1537: attaching collars $\torus^2\times(0,1)$ to a pseudotorus again
1538: gives $\torus^2\times(0,1)$.
1539: 
1540: We conclude the following from this.
1541: 
1542: \begin{lemma}
1543: For almost all $c$ each connected component
1544: of $N_c$ will be homeomorphic to $\torus^3$. The exceptions are those
1545: $c$ that are critical values of the function $\overline f$ on $\torus^2$.
1546: \end{lemma}
1547: 
1548: For a generic $f$ we obtain a generic Morse function $\overline f$
1549: on $\torus^2$. For such a function, there must be an interval $[c_1,c_2]$
1550: such that, whenever we have $c\in[c_1,c_2]$,
1551: the level line $\overline f=c$ contains a closed curve non-homologous to
1552: zero in $\torus^2$. The preimage $N_c$ of this level line in $\torus^4$
1553: is a $3$-torus non-homologous to zero.
1554: Thus, we get the following.
1555: 
1556: \begin{lemma}
1557: There exist $c_1$, $c_2$ such that $c_1<c_2$ and both $N_{c_1}$
1558: and $N_{c_2}$ contain a connected component homeomorphic to $\torus^3$
1559: and non-homologous to zero.
1560: \end{lemma}
1561: 
1562: It remains to notice that whenever $c>c_1$ the hypersurface $N_{c_1}$
1563: is essentially below $M^3_c$, and whenever $c<c_2$ the hypersurface
1564: $N_{c_2}$ is essentially above $M^3_c$. Thus, for all $c$
1565: we have a non-homologous to zero $3$-torus which is either essentially
1566: above or essentially below $M^3_c$, and we are done in Case~2.
1567: 
1568: 
1569: 
1570: \begin{thebibliography}{99}
1571: \bibitem{N}
1572: S.~P.~Novikov. Hamiltonian formalism and a multivalued analog
1573: of Morse theory. (Russian) {\it Uspekhi Mat.\ Nauk}~{\bf37} (1982),
1574: no.~5, 3--49; translation in {\it Russian Math.\
1575: Surveys}~{\bf37}~(1982);
1576: {\tt http://genesis.mi.ras.ru/$\sim$snovikov/74.zip}.
1577: \bibitem{Z}
1578: A.\,V.\,Zorich. Novikov's problem on semiclassical motion of electron
1579: in homogeneous magnetic field close to rational. (Russian)
1580: {\it Uspekhi Mat.\ Nauk}~{\bf39}~(1984), no.\ 5, 235--236; translation
1581: in {\it Russian Math.\ Surveys}~{\bf39} (1984).
1582: \bibitem{N1}
1583: S.\,P.\,Novikov. Quasiperiodic structures in topology.
1584: {\it Topological methods in modern mathematics\/} (Stony Brook, NY, 1991),
1585: 223--233, Publish or Perish, Houston, TX, 1993.
1586: \bibitem{D}
1587: I.\,A.\,Dynnikov.
1588: A proof of S.\,P.\,Novikov's conjecture for the case of small
1589: perturbations of rational magnetic fields. (Russian)
1590: {\it Uspekhi Mat.\ Nauk\/} {\bf47} (1992), no. 3(285), 161--162;
1591: translation in {\it Russian Math. Surveys\/} {\bf47} (1992), no. 3,
1592: 172--173.
1593: \bibitem{D'}
1594: I.\,A.\,Dynnikov.
1595: S.\,P.\,Novikov's problem on the semiclassical motion of an electron.
1596: (Russian) {\it Uspekhi Mat. Nauk\/} {\bf48} (1993),
1597: no. 2(290), 179--180; translation in {\it Russian Math. Surveys\/}
1598: {\bf48} (1993), no. 2, 173--174.
1599: \bibitem{D1}
1600: I.\,A.\,Dynnikov.
1601: A proof of the conjecture of S.\,P.\,Novikov on the semiclassical
1602: motion of an electron. (Russian) {\it Mat.\ Zametki\/} {\bf53}
1603: (1993), no.~5, 57--68; translation in {\it Math Notes\/}
1604: {\bf53} (1993), no.~5--6, 495--501.
1605: \bibitem{NM}
1606: S.\,P.\,Novikov and A.\,Ya.\,Maltsev.
1607: Topological quantum characteristics observed in the investigation
1608: of the conductivity in normal metals. (Russian)
1609: {\it Pis'ma Zh.\ Eksp.\ Teor.\ Fiz.} {\bf63} (1996), no.~10, 809--813;
1610: translation in
1611: {\it JETP Letters\/} {\bf63} (1996), no.~10, 855--860.
1612: \bibitem{NM1}
1613: S.\,P.\,Novikov and A.\,Ya.\,Maltsev.
1614: Topological phenomena in normal metals. (Russian)
1615: {\it Uspekhi Phys.\ Nauk\/} {\bf41} (1998), no.~3, 231--239;
1616: {\tt arXiv:cond-mat/9709007}.
1617: \bibitem{LAK}
1618: I.\,Lifshitz, M.\,Ya.\,Azbel, M.\,I.\,Kaganov.
1619: {\it Elektronnaya teoriya metallov}. (Russian) Nauka, Moscow,
1620: 1971; translation: {\it Electron theory of metals}, Consultants
1621: Bureau, New York,  1973.
1622: \bibitem{N2}
1623: S.\,P.\,Novikov.
1624: Levels of quasiperiodic functions on a plane, and Hamiltonian systems.
1625: (Russian) {\it Uspekhi Mat.\ Nauk\/} {\bf54} (1999),
1626: no.~5(329), 147--148; translation in {\it Russian Math.\ Surveys\/}
1627: {\bf54} (1999), no.~5, 1031--1032;
1628: {\tt arXiv:math-ph/9909032}.
1629: \bibitem{M}
1630: A.~Ya.Maltsev.
1631: Quasiperiodic functions theory and the
1632: superlattice potentials for a two-dimensional electron gas.
1633: {\it Journal of Mathematical Physics\/} {\bf45}, no.~3 (March 2004),
1634: 1128--1149;
1635: {\tt arXiv:cond-mat/0302014}.
1636: \bibitem{DKN}
1637: V.~I.~Arnold, S.~P.~Novikov (editors), {\it Encyclopedia
1638: Mathematical Sciences, Dynamical Systems, vol 4: Completely
1639: Integrable Systems and Symplectic Geometry}. Springer Verlag,
1640: second edition (revised), 2001:\\
1641:  1.\ V.\,Arnold, A.\,Givental;
1642: \\
1643:  2.\ B.\,Dubrovin, I.\,Krichever, S.\,Novikov.
1644: \bibitem{Pi}
1645: T.\,Q.\,T.\,Le, S.\,Piunikhin, and V.\,Sadov.
1646: The geometry of quasicrystals. (Russian) {\it Uspekhi Mat.\
1647: Nauk\/} {\bf48} (1993), no.~1(289), 41--102; translation in {\it
1648: Russian Math.\ Surveys\/} {\bf48} (1993), no.~1, 37--100.
1649: \bibitem{LP}
1650: I.\,M.\,Lifshitz, V.\,G.\,Peschanski.
1651: Halvanomagnetic characteristics of metals with open Fermi surfaces,~II.
1652: (Russian) {\it Zh.\ Eksp.\ Teor.\ Fiz.} {\bf 38} (1960), 188--193;
1653: translation in {\it JETP.}.
1654: \bibitem{Ab}
1655: A.~A.~Abrikosov.
1656: {\it Osnovy torii metallov}. (Russian) Nauka, Moscow, 1987;
1657: translation: {\it Fundamentals of theory of metals},
1658: North-Holland, Amsterdam, 1988.
1659: \bibitem{Ki}
1660: C.\,Kittel, {\it Quantum theory of solids}. John
1661: Wiley \& Sons Inc., New York, London, 1963.
1662: \bibitem{Zi}
1663: J.\,M.\,Ziman. {\it Principles of the theory of solids}.
1664: Cambridge Univ.\ Press, 1972.
1665: \bibitem{MN}
1666: A.\,Ya.\,Maltsev and S.\,P.\,Novikov.
1667: Dynamical Systems, Topology, and Conductivity in Normal Metals.
1668: {\it Journal of Statistical Physics\/} {\bf115} no.~1, April 2004, 31-46;
1669: {\tt arXiv:cond-mat/0312708}.
1670: \bibitem{MN1}
1671: A.\,Ya.\,Maltsev and S.\,P.\,Novikov.
1672: Quasiperiodic functions and dynamical systems in quantum solid state
1673: physics. Dedicated to the 50th anniversary of IMPA. {\it Bull.\ Braz.\
1674: Math.\ Soc. (N.S.)\/} {\bf 34} (2003), no.~1, 171--210;
1675: {\tt arXiv:math-ph/0301033}.
1676: \bibitem{MN2}
1677: A.\,Ya.\,Maltsev, S.\,P.\,Novikov.
1678: Topology, Quasiperiodic functions and the transport phenomena.
1679: {\it Topology in condensed matters}, M.\,I.\,Monastyrsky (ed.),
1680: Sprinter Verlag, 2004, to appear;
1681: {\tt arXiv:cond-mat/0312710}.
1682: \bibitem{D2}
1683: I.\,A.\,Dynnikov.
1684: Semiclassical motion of the electron. A proof of the Novikov conjecture
1685: in general position and counterexamples.
1686: {\it Solitons, geometry, and topology\/{\rm:} on the crossroad},
1687: 45--73, Amer.\ Math.\ Soc.\ Transl.\ Ser.~2, 179,
1688: {\it Amer.\ Math.\ Soc.}, {\it Providence}, {\it RI}, 1997
1689: \bibitem{D3}
1690: I.\,A.\,Dynnikov.
1691: The geometry of stability regiones in Novikov's problem on
1692: the semiclassical motion of the electron. (Russian)
1693: {\it Uspekhi Mat.\ Nauk}, {\bf 54} (1999), no.~1, 21--60.
1694: \bibitem{RdL}
1695: R.\,De Leo. Numerical Analysis of the Novikov Problem
1696: of a Normal Metal in a Strong Magnetic
1697: Field. {\it SIAM Journal of Applied Dynamical Systems\/}
1698: {\bf 2}:4 (2003), 517--545.
1699: \bibitem{DM}
1700: I.\,A.\,Dynnikov and A.\,Ya.\,Maltsev.
1701: Topological characteristics of electron spectra in monocrystals.
1702: (Russian) {\it Zh.\ Eksp.\ Teor.\ Fiz.} {\bf112}:1
1703: (1997), 371--378; translation in {\it JETP\/} {\bf85}:1 (1997),
1704: 205--208.
1705: \bibitem{B}
1706: C.\,W.\,J.\,Beenaker.  Guiding-center-drift resonance in a
1707: periodically modulated two-dimensional electron gas.
1708: {\it Phys.\ Rev.\ Lett.} {\bf62} (1989), no.~17, 2020--2023.
1709: 
1710: 
1711: \end{thebibliography}
1712: 
1713: \end{document}
1714: