Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

On Brillouin Zones

1998
We then use these ideas to discuss focusing of geodesics in orbifolds of constant curvature. In the particular case of the Riemann surfaces H^2/Gamma(k), (k=2,3, or 5), we explicitly count the number of geodesics of length t that connect the point i to itself....Read more
arXiv:math/9806154v1 [math.MG] 29 Jun 1998 ON BRILLOUIN ZONES J. J. P. Veerman veerman@dmat.ufpe.br Physics Department UFPE Recife, Brazil M. M. Peixoto peixoto@impa.br IMPA Rio de Janeiro, Brazil A. C. Rocha acr@dmat.ufpe.br Mathematics Dept. UFPE Recife, Brazil S. Sutherland scott@math.sunysb.edu Mathematics Dept. SUNY Stony Brook, NY, USA Stony Brook IMS Preprint #1998/7 June 1998 Abstract Brillouin zones were introduced by Brillouin [Br] in the thirties to describe quantum mechanical properties of crystals, that is, in a lattice in R n . They play an important role in solid-state physics. It was shown by Bieberbach [Bi] that Brillouin zones tile the underlying space and that each zone has the same area. We generalize the notion of Brillouin Zones to apply to an arbitrary discrete set in a proper metric space, and show that analogs of Bieberbach’s results hold in this context. We then use these ideas to discuss focusing of geodesics in orbifolds of constant curvature. In the particular case of the Riemann surfaces H 2 /Γ(k)(k =2, 3, or 5), we explicitly count the number of geodesics of length t that connect the point i to itself. 1 Introduction In solid-state physics, the notion of Brillouin zones is used to describe the behavior of an electron in a perfect crystal. In a crystal, the atoms are often arranged in a lattice; for example, in NaCl, the sodium and chlorine atoms are arranged along the points of the simple cubic lattice Z 3 . If we pick a specific atom and call it the origin, its first Brillouin zone consists of the points in R 3 which are closer to the origin than to any other element of the lattice. This same zone can be constructed as follows: for each element a in the lattice, let L 0a be the perpendicular bisecting plane of the line between 0 and a (this plane is called a Bragg plane). The volume about the origin enclosed by these intersecting planes is the first Brillouin zone, b 1 (0). This construction also allows us to define the higher Brillouin zones as well: a point x is in b n if the line connecting it to the origin crosses exactly n 1 planes L 0a , counted with multiplicity. This notion was introduced by Brillouin in the 1930s ([Br]), and plays an important role in solid-state theory (see, for example, [AM] or [Jo2], also [Ti]). The construction which gives rise to Brillouin zones is not limited to consideration of crystals, however. For example, in computational geometry, the notion of the Voronoi cell corresponds exactly to the first Brillouin zone described above (see [PS]). We shall also see below how, after suitable generalization, this construction coincides with the Dirichlet domain of Riemannian geometry, and in many cases, to the focal decomposition introduced in [Pe1] (see also [Pe3]).
2 Veerman, Peixoto, Rocha, and Sutherland 1 2 2 2 2 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 4 4 Figure 1.1: On the right are the Brillouin zones for the lattice Z 2 in R 2 . On the left is the outer boundary of the third Brillouin zone for the lattice Z 3 in R 3 . With some slight hypotheses (see section 2), we generalize the construction of Brillouin zones to any discrete set S in a path-connected, proper metric space X. We generalize the Bragg planes above as mediatrices, defined here. Definition 1.1 For a and b distinct points in S , define the mediatrix (also called the equidistant set or bisector) L ab of a and b as: L ab = {x X d(x, a)= d(x, b)} . Now choose a preferred point x 0 in S , and consider the collection of mediatrices {L x 0 ,s } sS . These partition X into Brillouin zones as above: roughly, the n-th Brillouin zone B n (x 0 ) con- sists of those points in X which are accessible from x 0 by crossing exactly n 1 mediatrices. (There is some difficulty accounting for multiple crossings— see definition 2.7 for a precise statement.) One basic property of the zones B n is that they tile the space X: x i S B n (x i )= X and B n (x 0 ) B n (x 1 ) is small . Here, with some extra hypotheses, “small” means of measure zero. Furthermore, again with some extra hypothesis, each zone B n has the same area. (This property was “obvious” to Brillouin). Both results were proved by Bieberbach in [Bi] in the case of a lattice in R 2 . Indeed, he proves (as we do) that each zone forms a fundamental set for the group action of the lattice. His arguments rely heavily on planar Euclidean geometry, although he remarks that his considerations work equally well in R d and can be extended to “groups of motions in non-Euclidean spaces”. In [Jo1], Jones proves these results for lattices in R d , as well as giving asymptotics for both the distance from B n to the basepoint, and for the number of connected components of the interior of B n . In section 2, we show that the tiling result holds for arbitrary
arXiv:math/9806154v1 [math.MG] 29 Jun 1998 ON BRILLOUIN ZONES A. C. Rocha S. Sutherland J. J. P. Veerman M. M. Peixoto veerman@dmat.ufpe.br peixoto@impa.br acr@dmat.ufpe.br scott@math.sunysb.edu Physics Department UFPE Recife, Brazil IMPA Rio de Janeiro, Brazil Mathematics Dept. Mathematics Dept. SUNY UFPE Stony Brook, NY, USA Recife, Brazil Abstract Brillouin zones were introduced by Brillouin [Br] in the thirties to describe quantum mechanical properties of crystals, that is, in a lattice in Rn . They play an important role in solid-state physics. It was shown by Bieberbach [Bi] that Brillouin zones tile the underlying space and that each zone has the same area. We generalize the notion of Brillouin Zones to apply to an arbitrary discrete set in a proper metric space, and show that analogs of Bieberbach’s results hold in this context. We then use these ideas to discuss focusing of geodesics in orbifolds of constant curvature. In the particular case of the Riemann surfaces H2 /Γ(k) (k = 2, 3, or 5), we explicitly count the number of geodesics of length t that connect the point i to itself. 1 Introduction In solid-state physics, the notion of Brillouin zones is used to describe the behavior of an electron in a perfect crystal. In a crystal, the atoms are often arranged in a lattice; for example, in NaCl, the sodium and chlorine atoms are arranged along the points of the simple cubic lattice Z3 . If we pick a specific atom and call it the origin, its first Brillouin zone consists of the points in R3 which are closer to the origin than to any other element of the lattice. This same zone can be constructed as follows: for each element a in the lattice, let L0a be the perpendicular bisecting plane of the line between 0 and a (this plane is called a Bragg plane). The volume about the origin enclosed by these intersecting planes is the first Brillouin zone, b1 (0). This construction also allows us to define the higher Brillouin zones as well: a point x is in bn if the line connecting it to the origin crosses exactly n − 1 planes L0a , counted with multiplicity. This notion was introduced by Brillouin in the 1930s ([Br]), and plays an important role in solid-state theory (see, for example, [AM] or [Jo2], also [Ti]). The construction which gives rise to Brillouin zones is not limited to consideration of crystals, however. For example, in computational geometry, the notion of the Voronoi cell corresponds exactly to the first Brillouin zone described above (see [PS]). We shall also see below how, after suitable generalization, this construction coincides with the Dirichlet domain of Riemannian geometry, and in many cases, to the focal decomposition introduced in [Pe1] (see also [Pe3]). Stony Brook IMS Preprint #1998/7 June 1998 2 Veerman, Peixoto, Rocha, and Sutherland 4 3 4 3 2 4 4 3 3 4 4 2 2 4 4 1 3 3 4 2 3 4 4 3 4 Figure 1.1: On the right are the Brillouin zones for the lattice Z2 in R2 . On the left is the outer boundary of the third Brillouin zone for the lattice Z3 in R3 . With some slight hypotheses (see section 2), we generalize the construction of Brillouin zones to any discrete set S in a path-connected, proper metric space X. We generalize the Bragg planes above as mediatrices, defined here. Definition 1.1 For a and b distinct points in S, define the mediatrix (also called the equidistant set or bisector) Lab of a and b as: Lab = {x ∈ X d(x, a) = d(x, b)} . Now choose a preferred point x0 in S, and consider the collection of mediatrices {Lx0 ,s }s∈S . These partition X into Brillouin zones as above: roughly, the n-th Brillouin zone Bn (x0 ) consists of those points in X which are accessible from x0 by crossing exactly n − 1 mediatrices. (There is some difficulty accounting for multiple crossings— see definition 2.7 for a precise statement.) One basic property of the zones Bn is that they tile the space X: [ Bn (xi ) = X and Bn (x0 ) ∩ Bn (x1 ) is small . xi ∈S Here, with some extra hypotheses, “small” means of measure zero. Furthermore, again with some extra hypothesis, each zone Bn has the same area. (This property was “obvious” to Brillouin). Both results were proved by Bieberbach in [Bi] in the case of a lattice in R2 . Indeed, he proves (as we do) that each zone forms a fundamental set for the group action of the lattice. His arguments rely heavily on planar Euclidean geometry, although he remarks that his considerations work equally well in Rd and can be extended to “groups of motions in non-Euclidean spaces”. In [Jo1], Jones proves these results for lattices in Rd , as well as giving asymptotics for both the distance from Bn to the basepoint, and for the number of connected components of the interior of Bn . In section 2, we show that the tiling result holds for arbitrary 3 On Brillouin Zones discrete sets in a metric space. If the discrete set is generated by a group of isometries, we show that each Bn forms a fundamental set, and consequently all have the same area (see Prop. 2.11). We now discuss the relationship of Brillouin zones and focal decomposition of Riemannian manifolds. If x1 (t) and x2 (t) are two solutions of a second order differential equation with x1 (0) = x2 (0) and there is some T 6= 0 so that x1 (T ) = x2 (T ), then the trajectories x1 and x2 are said to focus at time T . One can ask how the number of trajectories which focus varies with the endpoint x(T )— this gives rise to the concept of a focal decomposition (originally called a sigma decomposition). This concept was introduced in [Pe1] and has important applications in physics, for example when computing the semiclassical quantization using the Feynman path integral method (see [Pe3]). There is also a connection with the arithmetic of positive definite quadratic forms (see [Pe2], [KP], and [Pe3]). Brillouin zones have a similar connection with arithmetic, as can be seen in section 4. More specifically, consider the two-point boundary problem ẍ = f (t, x, ẋ), x(t0 ) = x0 , x(t1 ) = x1 , x, t, ẋ, ẍ ∈ R. Associated with this equation, there is a partition of R4 into sets Σk , where (x0 , x1 , t0 , t1 ) is in Σk if there are exactly k solutions which connect (x0 , t0 ) to (x1 , t1 ). This partition is the focal decomposition with respect to the boundary value problem. In [PT], several explicit examples are worked out, in particular the fundamental example of the pendulum ẍ = − sin x. Also, using results of Hironaka ([Hi]) and Hardt ([Ha]), the possibility of a general, analytic theory was pointed out. In particular, under very general hypotheses, the focal decomposition yields an analytic Whitney stratification. Later, in [KP], the idea of focal decomposition was approached in the context of geodesics of a Riemannian manifold M (in addition to a reformulation of the main theorem of [PT]). Here, one takes a basepoint x0 in the manifold M : two geodesics γ1 and γ2 focus at some point y ∈ M if γ1 (T ) = y = γ2 (T ). This gives rise to a decomposition of the tangent space of M at x into regions where the same number of geodesics focus. In order to study focusing of geodesics on an orbifold (M, g) with metric g via Brillouin zones, we do the following. Choose a base-point p0 in M and construct the universal cover X, lifting p0 to a point x0 in X. Let γ be a smooth curve in M with initial point p0 and endpoint p. Lift γ to γ̃ in X with initial point x0 . Its endpoint will be some x ∈ π −1 (p). The metric g on M is lifted to a metric g̃ on X by setting g̃ = π ∗ g. Under the above conditions, the group G of deck transformations is discontinuous and so π −1 (p0 ) ⊂ X is a discrete set. One can ask how many geodesics of length t there are which start at p0 end in p, or translated to (X, γ̃), this becomes: How many mediatrices Lx0 ,s intersect at x, as s ranges over π −1 (p0 )? Notice that if the universal cover of M coincides with the tangent space T Mx , the focal decomposition of [KP] and that given by Brillouin zones will be the same. If the universal cover and the tangent space are homeomorphic (as is the case for a manifold of constant negative curvature), the two decompositions are not identical, but there is a clear correspondence. However, if the universal cover of the manifold is not homeomorphic to the tangent space at the base point, the focal decomposition and that given by constructing Brillouin zones in the 4 Veerman, Peixoto, Rocha, and Sutherland universal cover are completely different. For example, let M be Sn , and let x be any point in it. The focal decomposition with respect to x gives a collection of nested n − 1-spheres centered at x; on each of these infinitely many geodesics focus (each sphere is mapped by the exponential to either x or its antipodal point). Between the spheres are bands in which no focusing occurs. (See [Pe3]). However, using the construction outlined in the previous paragraph gives a very different result. Since Sn is simply connected, it is its own universal cover. There is only one point in our discrete set, and so the entire sphere Sn is in the first zone B1 . The organization of this paper is as follows. In section 2, we set up the general machinery we need, and prove the main theorems in the context of a discrete set S in a proper metric space. Section 3 explores this in the context of manifolds of constant curvature. The universal cover is Rn , Sn , or Hn , and the group G of deck transformations is a discrete group of isometries (see [doC]). The discrete set S is the orbit of a point not fixed by any element of G under this discontinuous group. It is easy to see that the mediatrices in this case are totally geodesic spaces. From the basic property explained above, one can deduce that the n-th Brillouin zone is a fundamental domain for the group G in X. In section 4, we calculate exactly the number of geodesics of length t that connect the origin to itself in two cases: the flat torus R2 /Z2 and the Riemann surfaces H2 /Γ(p), for p ∈ {2, 3, 5}. While these calculations could, of course, be done independent of our construction, the Brillouin zones help visualize the process. In the final section, we give a nontrivial example in the case of a non-Riemannian metric, and mention a connection to the question of how many integer solutions there are to the equation ak + bk = n, for fixed k. Acknowledgments: It is a pleasure to acknowledge useful conversations with Federico Bonetto, Johann Dupont, Irwin Kra, Bernie Maskit, John Milnor, Chi-Han Sah, and Duncan Sands. Part of this work was carried out while Peter Veerman was visiting the Center for Physics and Biology at Rockefeller University and the Mathematics Department at SUNY Stony Brook; the authors are grateful for the hospitality of these institutions. 2 Definitions and main results In this section, we prove that under very general conditions, Brillouin zones tile (as defined below) the space in which they are defined, generalizing an old result of Bieberbach [Bi]. With stronger assumptions, we prove that these tiles are in fact well-behaved sets: they are equal to the closure of their interior. Notation: Throughout this paper, we shall assume X is a path connected, proper (see below) metric space (with metric d(·, ·)). We will make use the following notation: • Write an open r-neighborhood of a point x0 as Nr (x0 ) = {x ∈ X d(x0 , x) < r}. • Define the circumference as Cr (x0 ) = {x ∈ X d(x0 , x) = r}. 5 On Brillouin Zones • Their union is the closed disk of radius r, denoted by Dr (x0 ) = {x ∈ X d(x0 , x) ≤ r}. Definition 2.1 A metric space X is proper if the distance function d(x, ·) is a proper map for every fixed x ∈ X. In particular, for every x ∈ X and r > 0, the closed ball Dr (x) is compact. Such a metric space is also sometimes called a geometry (See [Ca]). Note if X is proper, it is locally compact and complete. The converse also holds if X is a geodesic metric space (see Thm. 1.10 of [Gr]). The metric spaces considered here need not be geodesic. Definition 2.2 The space X is called metrically consistent if, for all x in X, R > r > 0 in R, and for each a ∈ CR (x), there is a z ∈ Cr (x) satisfying Nd(z,a) (z) ⊆ NR (x) and Cd(z,a) (z) ∩ CR (x) = {a}. This property is satisfied for any Riemannian metric. Note that any mediatrix La,b separates X, that is: X − Lab contains at least two components (one containing the point a and the other b). Definition 2.3 We say that the mediatrix Lab is minimally separating if for any subset L̃ ⊂ Lab with L̃ 6= Lab , the set X − L̃ has one component. If a mediatrix L it is minimally separating, then X − L has exactly two components. Note that if a separating set L ⊆ X contains a non-empty open set V , it cannot be minimally separating. For if A and B are disjoint open sets containing X − L, then so are à = A − (A ∩ V ) and B̃ = B − (B ∩ V ). Now let x be any point in V . Then it is easy to see that the disjoint open sets à ∪ V and B̃ cover X − (L − x). Thus L − x separates X, so L could not have been minimal. We define the following sets L− 0a = {x ∈ X d(0, x) − d(a, x) < 0} , L+ 0a = {x ∈ X d(0, x) − d(a, x) > 0} . Definition 2.4 Two minimally separating sets L0a and L0b are topologically transversal if they + are disjoint, or if for each x ∈ L0a ∩ L0b and every neighborhood V of x, the sets L+ 0a ∩ L0b ∩ V , + − − + − − L0a ∩ L0b ∩ V , L0a ∩ L0b ∩ V , L0a ∩ L0b ∩ V are all nonempty. Usually, there will be a discrete set of points S = {xi }i∈I in X which will be of interest. By discrete we mean that any compact subset of X contains finitely many points of S. Note that if lim inf d(a, b) > 0, then S is discrete. a,b∈S Definition 2.5 We say a proper, path connected metric space X is Brillouin if it satisfies the following conditions: 1: X is metrically consistent. 2: For all a, b in X, the mediatrices Lab are minimally separating sets. 6 Veerman, Peixoto, Rocha, and Sutherland 3: For any three distinct points 0, a, and b in X, the mediatrices L0a and L0b are topologically transversal. The last two conditions in the above definition may be weakened to apply only to those mediatrices Lab where a and b in S. In this case, we will say that X is Brillouin over S, if it is not obvious from the context. Figure 2.1: The set L(0,0),(a,a) contains two quarter-planes. 2 L√ 0a for R with the Manhattan Figure 2.2: L(0,0),(4,6) (thin solid line) and Figure 2.3: The mediatrices  L(0,0),(2,4) (thick grey line) are not transverse. metric and a in the lattice (m, n 2) . Example 2.6 Equip R2 with the “Manhattan metric”, that√is, d(p, q) = |p1 − q1 | + |p2 − q2 |. In this metric, a circle Cr (p) is a diamond of side length r 2 centered at p, so condition 1 is satisfied. However, condition 2 fails: if the coordinates of a point a are equal, then L0a consists of a line segment and two quarter-planes (see Fig. 2.1). If the discrete set S contains no such points, we can still run into trouble with topological transversality. For example, the mediatrices L(0,0),(2,4) and L(0,0),(4,6) both contain the ray {(t, 1) t ≥ 4} (Fig. 2.2). But, if we are careful, we can ensure that the space is Brillouin over S. To achieve this, if (0, 0) is the basepoint, we must have that for all pairs (a1 , a2 ) and (b√ 1 , b2 ) in S, a1 − a2 6= b1 − b2 . For example, take S to be an irrational lattice such as (m, n 2) m, n ∈ Z . It is easy to check that for all a, b ∈ S, the properties of definition 2.5 are true. (From this example, we see that to do well in Manhattan, one should be carefully irrational.) As mentioned in the introduction, for each x0 ∈ S, the mediatrices Lx0 a give a partition of X. Informally, those elements of the partition which are reached by crossing n−1 mediatrices from x0 form the n-th Brillouin zone, Bn (x0 ). This definition is impractical, in part because 7 On Brillouin Zones a path may cross several mediatrices simultaneously, or the same mediatrix more than once. Instead, we will use a definition given in terms of the number of elements of S which are nearest to x. This definition is equivalent to the informal one when X is Brillouin over S (see Prop. 2.13 below). We use the notation #(S) to denote the cardinality of the set S. x0 x0 Figure 2.4: Here we illustrate the definition of the sets bn (x0 ) and Bn (x0 ) for the lattice Z2 in R2 . In both pictures, the circle Cd(x,x0 ) (x) is drawn, and the basepoint x0 lies in the center of the square at the lower left. On the left side, the point x (marked by a small cross) lies in b5 , and # (Nr (x) ∩ S) = 4, while x0 is the only point of S on the circle. On the right, we have m = 4 and ℓ = 8, so x lies in all of the sets B5 , B6 , . . . , B12 . Definition 2.7 Let x ∈ X, let n be a positive integer, n ≤ #(S), and let r = d(x, x0 ). Then define the sets bn (x0 ) and Bn (x0 ) as follows. • x ∈ bn (x0 ) ⇐⇒ # (Nr (x) ∩ S) = n − 1 • x ∈ Bn (x0 ) ⇐⇒ # (Nr (x) ∩ S) = m with m + 1 ≤ n ≤ m + ℓ. and and Cr (x) ∩ S = {x0 }. # (Cr (x) ∩ S) = ℓ, where l, m ∈ Z+ Here the point x0 is called the base point, and the set Bn (x0 ) is the n-th Brillouin zone with base point x0 . Note that in the second part, if m = n − 1 and ℓ = 1, then x ∈ bn (x0 ). So bn (x0 ) ⊆ Bn (x0 ). Note also that the complement of bn (x0 ) in Bn (x0 ) consists of subsets of mediatrices (see Def. 1.1). Note also that bn (x0 ) is open and that Bn (x0 ) is closed. Finally, observe that for fixed x0 the sets bn (x0 ) are disjoint, but the sets Bn (x0 ) are not. In what follows it will be proved that {Bn (x0 )}n>0 cover the space X. Thus we can assign to each point x its Brillouin index as the the largest n for which x ∈ Bn (x0 ). This definition was first given in [Pe1]. The following lemma, which follows immediately from Def. 2.7, explains a basic feature of the zones, namely that they are concentric in in a weak sense. This property is also apparent from the figures. 8 Veerman, Peixoto, Rocha, and Sutherland Lemma 2.8 Any continuous path from x0 to Bn (x0 ) intersects Bn−1 (x0 ). The Brillouin zones actually form a covering of X by non-overlapping closed sets in various ways. This is proved in parts. The next two results assert that the zones B cover X, but the zones b do not. The first of these is an immediate consequence of the definitions. The second is more surprising and leads to corollary 3.4. The fact that the Bi (xn ) is the closure of bi (xn ) (and thus that the interiors do not overlap) is proved in proposition 2.12. We will use the word “tiling” for a covering by non-overlapping closed sets. Lemma 2.9 For fixed n the Brillouin zones tile X in the following sense: [ Bi (xn ) = X and bi (xn ) ∩ bj (xn ) = ∅ if i 6= j. i Figure 2.5: This example illustrates Lemma 2.9 and Thm. 2.10. Let S be the discrete set {(m, 0)} ∪ {(0, n)} , m, n ∈ Z in the Euclidean plane. On the left is the tiling given by Bi (0, 0) and in the middle is the tiling by Bi (2, 0). In both cases, b2 is shaded. On the right is the tiling given by B2 (xi ) as in Thm. 2.10. The sets b2 (0, 0), b2 (1, 0), and b2 (2, 0) have been shaded. Note that this S does not correspond to a group, nor does it satisfy the hypotheses of Prop. 2.11, because there are no isometries which permute S and do not fix the origin. Theorem 2.10 Let X be a proper, path connected metric space let S = {xi }i∈I be a discrete set. Then, for fixed n ≤ #(S), the sets {Bn0 (xi )}i∈I tile X in the following sense: [ Bn (xi ) = X and bn (xi ) ∩ bn (xj ) = ∅ if i 6= j. i Proof: First, we show that for any fixed n > 0 and each x ∈ X, there is an xi ∈ S with x ∈ Bn (xi ). Re-index S so that if S = {x1 , x2 , x3 , . . .} and i < j, then d(x, xi ) ≤ d(x, xj ). This can be done; since S is a discrete subset and closed balls Dc (xi ) are compact, the subsets of S with d(x, xi ) ≤ c are all finite. Let ri = d(x, xi ). We will show that x ∈ Bn (xn ). Note that rn ≥ rn−1 . Suppose first that rn > rn−1 , then Nrn (x) ∩ S contains exactly n − 1 points, and xn ∈ Crn (x) ∩ S. Thus x ∈ Bn (xn ). Note that if rn+1 > rn , then we would have x ∈ bn (xn ) ⊂ Bn (xn ). If, on the other hand, rn = rn−1 , then there is a k > 0 so that rn = rn−1 = . . . = rn−k , and so # (Nrn (x) ∩ S) = n − k − 1 ≤ n − 1. But then # (Crn (x) ∩ S) ≥ k + 1, and hence x ∈ Bn (xn ) as desired. 9 On Brillouin Zones For the second part, we show that bn (xi )∩bn (xj ) = ∅. If not, then there is a point x in their intersection. If ri = rj , then xi = xj , because by the definition of bn (xk ), {xk } = Crk (x) ∩ S. If not, then ri < rj . In this case, xi ∈ Dri (x) ⊂ Nrj (x) . Thus, since # (Nri (x) ∩ S) = n − 1, Nrj (x) must contain at least n points of S, a contradiction.  The next result indicates how this notion of tiling is related to the notion of fundamental domain. Proposition 2.11 Let S be a discrete set in a metric space X as in theorem 2.10. Suppose that for each xi in S there is an isometry gi of X such that gi (x0 ) = xi , gi permutes S and the only gi which leaves x0 fixed is the identity. Then there is a set F (the fundamental set), satisfying: bn (x0 ) ⊆ [ i gi (F ) = X F and ⊆ Bn (x0 ) with gi (F ) ∩ gj (F ) = ∅ (i 6= j). Proof: Suppose that x ∈ bn (x0 ). From definition 2.7 and the fact that the gi are isometries, we see that this is equivalent to gi (x) ∈ bn (xi ). Thus gi (bn (x0 )) = bn (xi ). Now apply theorem 2.10. A similar reasoning proves the statement for Bn (x0 ).  Remark: The fundamental set F is not necessarily connected. Also, note that it follows from this proposition that Bi (x0 ) is scissors congruent to Bj (x0 ) (see [Sah] for a discussion of scissors congruence). In particular, this implies immediately that the Bi all have the same area. Note that this result does not hold if S is not generated by a group of isometries. For example, the “face-centered cubic” structures common to many crystals do not satisfy this, and the corresponding Bn have different volumes, although they still tile the space. See also figure 2.5. The preceding results are valid without the transversality conditions on the metric, as the reader can check. We will now use the additional properties of the metric to derive a regularity result. Proposition 2.12 If X is Brillouin (over S), then Bn (xi ) is the closure of bn (xi ). Proof: Let x ∈ Bn (x0 ). If x ∈ bn (x0 ), we are done. So suppose x 6∈ bn (x0 ), and then by Def. 2.7 (with r = d(x, x0 )), Nr (x) contains m ≤ n − 1 points of S and Cr (x) contains ℓ ≥ n − m with ℓ ≥ 2 points of S, including x0 . Thus x lies in the intersection of ℓ − 1 ≥ 1 ℓ−1 mediatrices {Lx0 ,xi }i=1 . Suppose ℓ = 2, so x lies on a single mediatrix L1 . Then by condition 2, x is in the + − + ′ ′ ′ closure of L− 1 and L1 . Thus if x ∈ L1 and close enough to x, then x ∈ bm+1 (x0 ). If x ∈ L1 and sufficiently close to x, then x′ ∈ bm+2 (x0 ). From our choices, we see that m = n − 1 or m = n − 2. In either case, we are done. The proof for ℓ > 2 is slightly more involved (we have to use transversality), but the idea is the same. The point x lies in the intersection of at least two mediatrices. Essentially, we 10 Veerman, Peixoto, Rocha, and Sutherland need show that the mediatrices disentangle in every neighborhood of x. We will prove that for all j ∈ {1, . . . , ℓ}, x ∈ bm+j (x0 ). Note that n − ℓ ≤ m ≤ n − 1, and so showing this is sufficient. Again, since mediatrices are minimally separating, we have that for all i ∈ {1, . . . , ℓ − 1} + there exist open sets L− i and Li in any neighborhood of x for which d(x, x0 ) < d(x, xi ) for x ∈ L− i and d(x, x0 ) > d(x, xi ) for x ∈ L+ i . Noting that xi ∈ Cr (x), then from condition 1 of Def. 2.7, we have that for any small ρT> 0, there is a point z − ∈ Cρ (x) for which d(z − , x0 ) < d(z − , xi ) for all i ∈ {1, . . . , ℓ}. Thus i L− i is nonempty, and is contained in bm+1 (x0 ). By the same principle, we can choose a point T z + ∈ Cr (x) for which d(z + , x0 ) > d(z + , xi ). As a consequence, i L+ i is nonempty, and is a subset of bm+ℓ (x0 ). Now, bm+1 (x0 ) and bm+ℓ (x0 ) are disjoint. If ℓ > 1, their complement in a neighborhood of x must contain at least two mediatrices. By transversality (condition 3), these cannot coincide. Thus V must be intersected by some bm+i (x0 ) for i ∈ {1, . . . ℓ}. By induction one finishes the argument.  Remark: The references to mediatrices in the above proof may make it appear that we are relying on the informal definition of Bn rather than Definition 2.7. In fact, we are only using the fact that Bn \ bn consists of subsets of mediatrices, which was noted following Def. 2.7. In fact, for Brillouin spaces, the informal definition of bn by mediatrices and Definition 2.7 are equivalent, as the following proposition shows. Proposition 2.13 Let X be Brillouin over S. Then there is a path γ from x to x0 which crosses exactly n − 1 mediatrices with no multiple crossings if and only if x ∈ bn (x0 ). By “no multiple crossings” we mean that if Li and Lj are distinct mediatrices, γ intersects Li in at most a single point, and γ ∩ Li 6= γ ∩ Lj . Proof: First, suppose there is such a path γ from x0 to x. As long as γ(t) crosses no mediatrices, the number of elements of S contained in Dr (γ(t)) remains constant. If this number is i, then γ(t) ∈ bi (x0 ). Each time γ(t) crosses a single mediatrix, this number increases by 1. Thus, after crossing n − 1 mediatrices, γ(t) will be in bn (x0 ). Now, let x ∈ bn (x0 ). We must construct a path from x to x0 which crosses exactly n − 1 mediatrices. Let b∗n denote the connected component of bn (x0 ) which contains x, and choose a point y in Bn−1 (x0 ) ∩ ∂b∗n . Such a point always exists by Lemma 2.8 and Prop. 2.12. In fact, because ∂bn consists of subsets of mediatrices, topological transversality of mediatrices allows us to choose y so that it lies on a single mediatrix. There certainly is a path connecting x to y which crosses no mediatrices; by similar reasoning, we can choose x′ ∈ bn−1 (x0 ) and a path from x′ to y which crosses no mediatrices. We have now constructed a path from x ∈ bn to x′ ∈ bn−1 which crosses exactly one mediatrix— repeating the argument n − 2 times finishes the proof.  On Brillouin Zones 3 11 Brillouin zones in spaces of constant curvature In this section X will be assumed to be one of Rn , Sn , or Hn , all equipped with the standard metric, and let G be a discontinuous group of isometries of X. Denote the quotient X/G with the induced metric by (M, g). Then the construction of lifting to the universal cover, as outlined in the introduction, applies naturally to (M, g). In this section we describe focusing of geodesics in (M, g) by Brillouin zones in X. The discrete set S is given by the orbit of a chosen point in X (which we will call the origin) under the group of deck-transformations G. The fact that the Brillouin zones are fundamental domains is now a direct corollary of proposition 2.11. The regularity conditions of Def. 2.5 are easily verified in the present context. We do this first. Lemma 3.1 If X is either Rn , Sn , or Hn , then a mediatrix Lab in X is an (n−1)-dimensional, totally geodesic subspace consisting of one component, and X − Lab has two components. Proof: This is easy to see if we change coordinates by an isometry of X, putting a and b in a convenient position, say as x and −x. The mediatrix Lx,−x is easily seen to satisfy the conditions (in the case of Sn , it is the equator, and for the others, it is a hyperplane). The conclusion follows.  Proposition 3.2 All such spaces X are Brillouin (see definition 2.5). Proof: As remarked before, the first condition is satisfied for any Riemannian metric. The second condition is also easy. It suffices to observe that the subspaces of Lemma 3.1 are minimally separating. To prove the topological transversality condition, note that lemma 3.1 implies that if L0a and L0b coincide in some open set, then they are equal. One easily sees that then we must have L0a = L0b = Lab = L. Applying the second part of the lemma to L0a , L0b , and Lab , we see that they must separate X in at least 3 components, namely one for each point a, b, and 0. This leads to a contradiction.  Remark: Note that in fact, the mediatrices are transversal in the usual sense. If L0a and L0b coincide in an open set, then their tangent spaces also coincide at some point. Uniqueness of solutions of second order differential equations then implies L0a = L0b . Recall that a metric space X is called rigid if the only isometry which fixes each point of a nonempty open subset of X is the identity. It is not hard to see that Sn , Hn , and Rn are rigid spaces. See [Ra] for more details of rigid metric spaces and for the proof of the following result. Recall that the stabilizer in G of a point x ∈ X consists of those elements of G that fix x. Proposition 3.3 Let G be a discontinuous group of isometries of a rigid metric space X. Then there exists a point y of X whose stabilizer Gy is consists of the identity. 12 Veerman, Peixoto, Rocha, and Sutherland We now return to Brillouin zones as defined in the last section. Recall that G is a group of isometries of X that acts discontinuously on points in X. Let x0 be a point in X whose stabilizer under the action of G is trivial. Let x ∈ X be a point in the orbit of x0 under G, that is to say, x = γ(x0 ) for some γ ∈ G − {e}, where e is the identity of G. Let [x0 , x] be a geodesic segment of minimal length whose endpoints are x0 and x. Then Bn (x0 ), the n-th Brillouin zone relative to x0 , is the set of points in X such that the segment [x0 , x] intercepts exactly n − 1 mediatrices Lx0 ,y , where y is in the orbit of x0 under the group G. Proposition 2.11 immediately implies the most important fact about Brillouin zones in this setting. Figure 3.1: Brillouin zones for P SL(2, Z) in the hyperbolic disk. We have transported the “usual” upper half-plane representation using the map z 7→ iz+1 . On the left is shown the sets Bn ( 4i ), which give fundamental z+i domains as in Cor. 3.4. On the right, 0 is taken as a basepoint. Since the origin has a non-trivial stabilizer, the corresponding Brillouin zones give a double cover of the fundamental domains. Corollary 3.4 Let X be Rn , Sn , or Hn , and let G be a discontinuous group of isometries of X. Let x0 ∈ X be such that its stabilizer Gx0 under G is trivial. Then for every positive integer n, the n-th Brillouin zone Bn (x0 ) is a fundamental domain for the action of G on points in X. Its boundary is the union of pieces of totally geodesic subspaces and equals the boundary of its interior. Remark: The first Brillouin zone B1 (0) is the usual Dirichlet fundamental domain for the action of G. Furthermore, even when Gx0 is not trivial, Bn (x0 ) is a k-fold cover of a fundamental domain. As pointed out in the introduction, the number of geodesics that focus in a certain point is counted in the lift. So if a given point x ∈ X is intersected by n mediatrices, it is reached by n + 1 geodesics of length d(0, x) emanating from the reference point (the origin). In the next section, we give more specific examples of this. Finally, we state a conjecture. 13 On Brillouin Zones Conjecture 3.5 Let (X, g̃) be the universal cover of a d-dimensional smooth Riemannian manifold (M, g) as described in the construction. For a generic metric g on M , no more than d mediatrices intersect in any given point y of X. This conjecture acquires perhaps even more interest (and certainly more structure), when one restricts the collection of metrics on M to conformal ones ([Mas]). A result in this direction for M = R2 /Z2 can be found in [Jo1]. 4 Focusing in two Riemannian examples In this section, we give two examples (one of them new as far as we know) of focusing. Suppose that at t = 0 geodesics start emanating in all possible directions from the a point. At certain times t1 , t2 , ...., we may see geodesics returning to that point. We derive expressions for the number of geodesics returning at tn in two cases. First, as introductory example we will discuss this for M = R2 /Z2 (a more complete discussion of this example can be found in [Pe3]). Second, we will deal with a much more unusual example, namely M = H2 /Γ(k), where Γ(k) is a subgroup of P SL(2, Z) called the principal congruence subgroup of level k (defined in more detail below). We note that it seems to be considerably harder to count geodesics that focus in points other than our basepoint. Before continuing, consider the classical problem of counting Rg (n), the number of solutions in Z2 of p2 + q 2 = n. Let n = 2α k Y pβi i i=1 ℓ Y qiγi j=1 be the prime decomposition of the number n, where pi ≡ 1(mod 4) and qi ≡ 3(mod 4). The following classical result of Gauss (see [NZM]) will be very useful. Lemma 4.1 Rg (n) is zero whenever n is not an integer, or any of the γi is odd. Otherwise, Rg (n) = 4 k Y (1 + βi ). i=1 Example 4.2 Choose an origin in M = R2 /Z2 and lift it to the origin in R2 . Our discrete set S is then Z2 . Let ρx (t) be the number of geodesics of length t that connect the origin to the point x ∈ M . Proposition 4.3 In the flat torus, the number of geodesics of length t that connect the the origin to itself is given by  Rg (t2 ), if t2 ∈ N ρ0 (t) = 0, otherwise. 14 Veerman, Peixoto, Rocha, and Sutherland Proof: Notice that by definition geodesics of length t leaving from the origin in R2 reach the points contained in Ct (0). Only if t2 is an integer does this circle intersect points in Z2 .  Example 4.4 We now turn to the next example. Recall that P SL(2, Z) can be identified with the group of two by two matrices with integer entries and determinant one, and with multiplication by −1 as equivalence. For each k, the group Γ(k) is the subgroup of P SL(2, Z) given by Γ(k) =  a b c d   ∈ P SL(2, Z) a ≡ d ≡ 1 (mod k), b ≡ c ≡ 0 (mod k), . This group has important applications in number theory. The action of Γ(k) on H2 is given by the Möbius transformations   az + b a b g(z) = where ∈ Γ(k). c d cz + d We point out that for k = 2, 3, or 5, the surface H2 /Γ(k) is a sphere with 3, 4, or 12 punctures (see [FK]). We will find it more convenient to work in the hyperbolic disk D2 , which is the universal cover of H2 /Γ(k). We shall choose a representation of Γ(k) in the disk so that i ∈ H2 corresponds to the origin. This will allow us to determine the focusing of the geodesics which emanate from i. Note that the surface H2 /Γ(k) has special symmetries with respect to i: for example, i is the unique point fixed by the order 2 element of P SL(2, Z). Figure 4.1: The orbit of i under Γ(2) transported to the hyperbolic disk, and the corresponding Brillouin zones. Each zone Bn forms a fundamental domain for a 3-punctured sphere. 15 On Brillouin Zones Lemma 4.5 The action of the fundamental group of the surface H2 /Γ(k) can be represented as    r − is p + iq r + p ≡ 1 (mod k), r − p ≡ 1 (mod k) 2 2 2 2 and p + q + 1 = r + s , p − iq r + is s + q ≡ 0 (mod k), s − q ≡ 0 (mod k) acting on D2 . We shall denote this particular representation as the group Γ⊙ (k). Proof: Following the conventions in [Be], define φ : D2 → H2 , φ(z) = i z+1 . −z + 1 Push back the transformation g ∈ Γ(k) from H2 to D2 by g → φ−1 gφ to obtain a representation of g ∈ Γ(k) as a transformation acting on D2 . The matrix representation of this transformation is given by:   a+d b−c a−d b+c 2 +i 2 2 −i 2 , Ag =  b+c a+d b−c a−d 2 +i 2 2 −i 2 where det Ag = 1, since this matrix is conjugate to g, whose determinant is equal to 1. Let p = (a − d)/2 r = (a + d)/2 q = −(b + c)/2 s = −(b − c)/2 and Ag now written as Ag =  r − is p + iq p − iq r + is  . Here the numbers p, q, r, s are in Z and must satisfy the following congruence conditions: r + p ≡ 1 (mod k), r − p ≡ 1 (mod k) s + q ≡ 0 (mod k), s − q ≡ 0 (mod k) Since the determinant of Ag is equal to 1, we must also have p2 + q 2 + 1 = r 2 + s 2 .  We need another auxiliary result before we state the main result of this section. Lemma 4.6 Let (p, q) and (r, s) be two points in Z2 such that the integers A = p2 + q 2 and B = r 2 + s2 are relatively prime, and let ϕ be a rotation. Now ϕ(p, q) = (p′ , q ′ ) and ϕ(r, s) = (r ′ , s′ ) are in Z2 if and only if ϕ is a rotation by a multiple of π/2. 16 Veerman, Peixoto, Rocha, and Sutherland Proof: Let c be the cosine of the angle of rotation. We have c= p′ p + q ′ q r ′ r + s′ s = . A B Thus if p′ p + q ′ q and r ′ r + s′ s are not both equal to zero, A p′ p + q ′ q = . ′ ′ rr+ss B Because A and B are relatively prime and surely |p′ p + q ′ q| is less than or equal to A, and similarly for B, we have that p′ p + q ′ q = ±A and r ′ r + s′ s = ±B. This implies the result.  Now we define a counter just as before. Choose a lift of M = D2 /Γ⊙ (k) so that 0 ∈ M lifts to 0 ∈ D2 . Let γx (t) be the number of geodesics of length t that connect the origin to the point x ∈ M . 512 512 256 256 128 128 64 64 32 32 16 16 8 8 4 4 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 0 1 2 3 4 5 6 7 8 Figure 4.2: The non-zero values of ρ0 (t) for t ≤ 25 (left) and γ0 (t) for t ≤ 8 (right), which count how many geodesics of length t connect the origin to itself in the R2 /Z2 and D2 /Γ⊙ (2), respectively. Theorem 4.7 In the surface H2 /Γ(k), the number of geodesics which connect the point i ∈ H2 to itself is given by 2 1 t − 1)Rg (cosh2 t) for k = 2 4 Rg (cosh   2 1 cosh2 t−1 Rg (cosh t) for k = 3 4 Rg 9   cosh2 t−1 1 Rg (cosh2 t) for k = 5 4 Rg 25 Note that in all cases, the number is nonzero only if cosh2 t ∈ N. 17 On Brillouin Zones Proof: We shall work in the disk, rather than in H2 . Let S be the orbit of 0 ∈ D2 under Γ⊙ (k). Then the number of such geodesics is exactly the number of distinct points of S which lie on on the circle Ct (0) of radius t and centered at the origin. If x ∈ S, then by Lemma 4.5, it is of the form p+iq r+is with p, q, r, s integers satisfying 2 2 2 2 p + q = r + s − 1. Let n be their common value, that is, n = p2 + q 2 = r 2 + s2 − 1. We will first count the number of 4-tuples (p, q, r, s) that solve this equation, momentarily ignoring the congruence conditions. Note that the point x has Euclidean distance to the origin given by |x|2e = p2 + q 2 n . = r 2 + s2 n+1 The hyperbolic length of the geodesic connecting the origin to this point is t = arctanh (|x|e ). q n , or, equivalently, n = cosh2 t − 1. This gives that γ0 (t) is only non-zero when t = arctanh n+1 To count the number of intersections of Ct (0) with S for these values of t, observe that we can use Gauss’ result to count the number of pairs (p, q) such that p2 + q 2 = n. This number is given by Rg (n). For each such pair (p, q), we have a number of choices to form x= p + iq r + is By the above, this number is equal to Rg (n + 1). Thus, γ0 (t) is at most Rg (n)Rg (n + 1). However, we have over-counted: some of our choices for p, q, r, s represent the same point x ∈ S, and some of them may not satisfy the congruence conditions, which we have so far ignored. We will first account for the multiple representations, and then account for the congruence relations. q n Let p, q, r, s ∈ Z be as above, giving a point x = p+iq at distance t = arctanh r+is n+1 from the origin. If we multiply the numerator and denominator of x by eiθ , then x will remain unchanged. Because of the requirement that p2 + q 2 = r 2 + s2 − 1 = n, this is the only invariant, and by Lemma 4.6, θ must be a multiple of π2 for the numerator and denominator to remain Gaussian integers. We see that in our counting, we have represented our point x in 4 different ways: −q + ip −p − iq q − ip p + iq = = = x= r + is −s + ir −r − is s − ir meaning we have over-counted by a factor of at least 4. Now we account for the congruence conditions. First, consider the case k = 2. Note that q + s ≡ 0 (mod 2) if and only if p + r ≡ 1 (mod 2), because p2 + q 2 + 1 = r 2 + s2 , so we need only check this one condition. If the representation p+iq r+is fails to satisfy our parity condition, then q+s ≡ 1 ( mod 2) and consequently 18 Veerman, Peixoto, Rocha, and Sutherland p + r ≡ 0 ( mod 2). This means that the representation −q+ip −s+ir of this same point does satisfy the 2 1 parity conditions, giving exactly 4 Rg (cosh t − 1)Rg (cosh2 t) distinct points of S at distance t from the origin. If k = 3, then since k is odd, the congruence conditions on p, q, r, and s imply that r ≡ 1(mod 3) and p ≡ q ≡ s ≡ 0(mod 3). Note that the equation p2 + q 2 = n and p ≡ q ≡ 0 (mod 3) will be satisfied exactly Rg (n/32 ) times. (Recall that if n/9 is not an integer, Rg (n/9) = 0.) For fixed n, let (p, q) be any one of the solutions. We need to decide how many solutions the equation r 2 + s2 = n + 1 with r ≡ 1 (mod 3) and s ≡ 0 (mod 3) admits. The solution of the first equation implies that 3 divides n. Thus r 2 + s2 ≡ 1 (mod 3). Consequently, we have 4 choices mod 3 for the pair (r, s), namely (0, 1), (1, 0), (0, 2), and (2, 0). Let (p, q, r, s) ∈ Z2 ×Z2 be any solution to n = p2 +q 2 = r 2 +s2 −1 with p ≡ q ≡ 0( mod 3). For each choice of (p, q), we have exactly Rg (n+1) choices of (r, s). Now let R denote the product of the rotations by π/2 on each of the components of Z2 × Z2 . Using Lemma 4.6, we see that all such solutions can be obtained from just one by applying R repeatedly. It is easy to check that each quadruple of solutions thus constructed runs exactly once through the above list. Since precisely one out of the four associated solutions is compatible with the conditions, the total number of solutions is exactly: 1 n Rg (n + 1). Rg 4 9 Using the relationship between the Euclidean distance and the Poincaré length as before gives the result. If k = 5, the proof for k = 3 can be literally transcribed to obtain the result.  Remark: Note that the above results do not hold if k is not one of the cases mentioned. The primary difficulty is that for prime k ≥ 7, there are solutions which are not related by applying the rotation R. However, the argument does give 14 Rg (cosh2 t − 1)/k2 Rg (cosh2 t) as an upper bound for H2 /Γ(k) when k is an odd prime. Note that the surface H2 /Γ(k) is of genus 0 if and only if k ≤ 5 (see [FK]). 5 Non-Riemannian examples The present context is certainly not restricted to Riemannian metrics. As an indicator of this we now discuss a different set of examples. 19 On Brillouin Zones Let k be a positive number greater than one. Equip R2 with the distance function 1/k  k x − y k= |x1 − y1 |k + |x2 − y2 |k and let the discrete set S be given by Z2 . For k not equal to 2, this is not a Riemannian metric, yet all conclusions of section 2 hold. In particular, each Brillouin zone forms a fundamental domain. Note that determining the zones by inspecting the picture requires close attention! 10 2 5 1 y 0 y 0 -1 -5 -2 -10 -2 -1 0 1 2 -10 -5 0 5 10 x x Brillouin zones for the lattice Z2 in R2 with the metric |x1 − y1 |4 + |x2 − y2 |4 Figure 2.3 and Example 2.6, which deal with the case k = 1, the “Manhattan metric”. Figure 5.1: 1/4 . See also Now the problem of determining Ct (0) ∩ S for any given t is unsolved for general k. In fact, even for certain integer values of k greater than 2, it is not known whether Ct (0) ∩ S ever contains at least two points that are not related by the symmetries of the problem. For k = 4, the smallest t for which Ct (0) ∩ S has at least two (unrelated) solutions is given by t4 = 1334 + 1344 = 1584 + 594 . However, for k ≥ 5, it unknown whether this can happen at all (see [SW]). There are some things that can be said, however. In the situation where k ∈ {3, 4, 5, . . .}, the mediatrices intersect the coordinate axes only in irrational points or in multiples of 1/2. For if x = (p/q, 0) is a point of a mediatrix L(0,0),(a1 ,a2 ) , we have |p|k = |p − qa1 |k + |a2 q|k (p 6= 0, q 6= 0) . By Fermat’s Last Theorem, this has no solution unless either p = qa1 or a2 = 0. In the first case, pq = ±a2 , which can only occur if the lattice point is of the form (a2 , ±a2 ). If a2 = 0, then p a1 q = 2 . In particular, there is no nontrivial focusing along the axes. 20 Veerman, Peixoto, Rocha, and Sutherland To compute Figure 5.1, we took advantage of the smoothness of the metric. Not all metrics are sufficiently smooth for this procedure to work. Even for Riemannian metrics, in general the distance function is only Lipschitz, which will not be sufficiently smooth. For each a = (a1 , a2 ) ∈ Z2 , define a Hamiltonian: Ha (x) =k x − a k − k x k . The mediatrix L0a corresponds to the level set Ha (x) = 0. Because Ha (x) is smooth, we have uniqueness of solutions to Hamilton’s equations. In the current situation, where the dimension is two, the level set consists of one orbit. Thus, one can produce the mediatrix by numerically tracing the zero energy orbits of the above Hamiltonian. As mentioned above, for a general Riemannian metric, the distance function is only Lipschitz. This means we have no guarantee that the solutions of the above differential equation are unique. Indeed, there are examples of multiply connected Riemannian manifolds with selfintersecting mediatrices, as will be shown in a forthcoming work. References [AM] N. W. Ashcroft, N. D. Mermin, Solid State Physics, Holt, Rhinehart, and Winston, 1976. [Be] A. F. Beardon, The Geometry of Discrete Groups, Springer-Verlag, 1983. [Bi] L. Bieberbach, Über die Inhaltsgleichheit der Brillouinschen Zonen, Monatshefte für Math. und Phys. 48 (1939), 509–515. [Br] L. Brillouin, Wave Propagation in Periodic Structures, Dover, 1953. [Ca] J. Cannon, The Theory of Negatively Curved Spaces and Groups, in: Ergodic Theory, Symbolic Dynamics, and Hyperbolic Spaces. (ed: T. Bedford, M. Keane, C. Series). Oxford University Press, 1991. [doC] M. do Carmo, Riemannian Geometry, Birkhäuser, 1992. [FK] H. Farkas and I. Kra, Automorphic Forms for Subgroups of the Modular Group. II: Groups Containing Congruence Subgroups, J. d’Analyse Math. 70 (1996), 91-156. [Gr] M. Gromov, Structures métriques pour les variétés riemanniennes. (rédigé par Lafontaine, J. et Pansu, P. Cedic/F. Nathan). 1981. Paris. [Ha] R. M. Hardt, Stratifications of real analytic mappings and images, Inv. Math. 28 (1975), 193–208. [Hi] H. Hironaka, Subanalytic sets. Number theory, in: Algebraic geometry and commutative algebra in honor of Y. Akizuki, Kinokuniya Publications, 1973, 453–493. On Brillouin Zones 21 [Jo1] G. A. Jones, Geometric and Asymptotic Properties of Brillouin Zones in Lattices, Bull. Lond. Math. Soc. 16 (1984), 241–263. [Jo2] H. Jones, The Theory of Brillouin Zones and Electronic States in Crystals, NorthHolland, Amsterdam, 1975. [KP] I. A. K. Kupka, M. M. Peixoto, On the Enumerative Geometry of Geodesics, in: From Topology to Computation, (ed: J. E. Marsden, M. Shub), Springer, 1993, 243–253. [Mas] B. Maskit, personal communication. [NZM] I. Niven, H. S. Zuckerman, H. L. Montgomery, An Introduction to the Theory of Numbers, fifth edition, John Wiley and Sons, 1991. [Pe1] M. M. Peixoto, On end point boundary value problems J. Differential Equations 44 (1982), 273–280. [Pe2] M. M. Peixoto, Sigma décomposition et arithmétique de quelqes formes quadratiques définies positives, in: R. Thom Festschift volume: Passion des Formes (ed: M. Porte), ENS Editions, 1994, 455–479. [Pe3] M. M. Peixoto, Focal Decomposition in Geometry, Arithmetic, and Physics, in: Geometry, Topology and Physics, (ed: Apanasov, Bradlow, Rodrigues, & Uhlenbeck), de Gruyter & Co., 1997, 213–231. [PT] M. M. Peixoto, R. Thom. Le point de vue énumératif dans les problèmes aux limites pour les équations différentielles ordinaires. I: Quelques exemples., C. R. Acad. Sc. Paris I 303 (1986), 629–632; Erratum, C. R. Acad. Sc. Paris I 307 (1988), 197–198; II: Le théoréme., C. R. Acad. Sc. Paris I 303 (1986), 693–698. [PS] F. P. Preparata, M. I. Shamos, Computational Geometry, Springer, 1985. [Ra] J. G. Ratcliffe, Foundations of Hyperbolic Manifolds, Springer, 1994. [Sah] C. H. Sah, Hilbert’s Third Problem: Scissors Congruence, Research Notes in Mathematics 33, Pitman, 1979. [SW] C. M. Skinner, T. D. Wooley, Sums of k-th Powers, J. Reine Angew. Math. 462 (1995), 57–68. [Ti] M. Tinkham, Group Theory and Quantum Mechanics, McGraw-Hill, 1964.