Packing theory derived from and products of linear forms

S. E. Graiff Zurita*1, B. Kane†2, and R. Oishi-Tomiyasu‡3

1Graduate School of Mathematics, Kyushu University 2Department of Mathematics, University of Hong Kong 3Institute of Mathematics for Industry (IMI), Kyushu University

Abstract

Parastichies are patterns observed in plants and numerical patterns generated using golden angle method. We generalize this method for botan- ical pattern formation, by using Markoff theory and the theory of product of linear forms, to obtain a theory for (local) packing of any Riemannian mani- folds of general dimensions n with a locally diagonalizable metric, including the Euclidean spaces. Our method is based on the property of some special lattices that the density of the lattice packing maintains a large value for any scale transfor- mations in the directions of the standard Euclidean axes, and utilizes maps that fulfill a system of partial differential equations. Using this method, we prove that it is possible to generate almost uniformly distributed point sets on any real analytic Riemann surfaces. The packing density is bounded below by approximately 0.7. A packing with logarithmic- and a 3D analogue of the Vogel spi- ral are obtained as a result. We also provide a method to construct (n + 1)- dimensional Riemannian manifolds with diagonal and constant-determinant metrics from n-dimensional manifolds with such a metric, which generally works for n = 1,2. The obtained manifolds have the self-similarity of bio- logical growth characterized by increasing size without changing shape.

arXiv:2106.12333v3 [math.NT] 26 Jul 2021 Keywords: aperiodic packing; Markoff spectra; Lagrange spectra; products of linear forms; diagonalizable metric; inviscid Burgers equation; biological growth; pattern formation; Vogel spiral; Doyle spiral

*[email protected][email protected][email protected] (corresponding author)

1 1 Introduction

In botany, spiral patterns observed in plants, such as sunflower heads, cacti, and pinecones, are called parastichies. Parastichies can also be observed in patterns that are numerically generated on a surface with circular symmetry using golden angle method. With this method, uniformly distributed point sets have been gen- erated on a cylindrical surface [10], a disk [30], [46], surfaces of revolution [38], sphere surface for mesh generation on globe [44], [21], and Poincare´ disc [40]. In particular, the cylinder case was studied by the Bravais brothers to provide a math- ematical model of the leaf arrangements on a stem (i.e., phyllotaxis). The disk case is known as the Vogel spiral (Figure 1). Introduction on this interdisciplinary topic for mathematicians are found in [6], [13] and [23]. [2] provides a more historical overview.

√ Figure 1: Left: Vogelspiral ne2πin/(1+γ1) (n > 0: integer) and images of the lattice shortest vectors (arrows) that indicate the directions of parastichies, Right: Doyle spiral of type (12,24). This article deals with a generalization which includes both cases.

In the golden angle method, a new point is generated for each 2πϕ ≈ 137.5◦- rotation as in (a) of Figure 2, where√ ϕ = 1/(1 + γ1) is the golden angle defined using the γ1 = (1+ 5)/2. As shown in (b) of Figure 2, the generated patterns can be regarded as the image of the lattice with the basis of Eq.(1).

1 ϕ  1 , , ϕ = , ε0 : constant. (1) 0 ε0 1 + γ1 Various variations of the Vogel spiral have been created by substituting dis- tinct ϕ [34]. The Doyle spiral (Figure 1) has also been addressed as a pattern of parastichies [5]. In this study, we generalize the golden angle method to generate uniformly distributed point sets on various surfaces without circular symmetry and higher- dimensional Riemannian manifolds. As a packing, we shall consider an arrange- ment of non-overlapping n-dimensional balls of the identical size. Although it is not the scope of this article to calculate the varying radius of each circle as in the

2 (a)4 (b)

6 1 5 6 5 4 3 4 3 2 2 2 O 1 1 3 The image of gives the coordinates 5 of leaves on a stem.

Figure 2: (a) Golden angle method, which generates a new point for every 2πϕ ≈ 137.5◦-rotation. (b) Interpretation of the golden angle method as a lattice point mapping (cf. [1]) case of circle packings including the Doyle spiral, the latter will be also mentioned as the case of conformal mapping (Example 2). If the set of the sphere centers is provided as the intersection of an open subset n of the Euclidean space R and a union of finitely many translations of a lattice, then the packing is called a periodic packing, and an aperiodic packing otherwise. If the set of the sphere centers is the intersection of the subset and a lattice, then it is n a lattice packing. For any full-rank lattice L ⊂ R , let ∆(L) be the packing density defined by Eq.(4). For any Bn ∈ GLn(R) and diagonal D ∈ GLn(R), let L(Bn) and L(DBn) be the lattices generated by the column vectors of Bn and DBn. In Section 2, the following problem will be discussed to generate aperiodic n packings on open subsets of R (or local charts of Riemannian manifolds) with the packing density fixed to a certain range. The lower bound for the packing density 0 is approximated as ∆n. 0 0 Problem 1 Determine the lattice basis Bn ∈ GLn(R) with ∆n = ∆ (L(Bn)). 0 0 ∆n := sup ∆ (L(Bn)), (2) Bn∈GLn(R), 0 ∆ (L(Bn)) := inf ∆(L(DBn)). D∈GLn(R): diagonal

0 Some L(Bn) actually attains the supremum value ∆n (Lemma 1). Problem 1 is reduced to the problem known as products of linear forms. In particular, the case of n = 2 can be better understood by using Markoff theory. The following will be deduced in (1) of Theorem 1, which enables us to expand the scope of the golden angle method to higher dimensions. √ π 3π ∆0 = √ ≈ 0.702, ∆0 = ≈ 0.389, 2 2 5 3 14 √ π2 5 5π2 ∆0 ≥ √ ≈ 0.183, ∆0 ≥ ≈ 0.076. 4 10 29 5 12 · 112

3 n N In addition to the lattices, a map f (x) from an open subset D ⊂ R to R (n ≤ N) with the Jacobian matrix J(x) satisfying (?) and (??), is used for our purpose. (?) tJJ is an invertible diagonal matrix for any x ∈ D.

(??) det tJJ = c2 for some constant c 6= 0.

N n For a fixed pair of a map f : D → R and a lattice L ⊂ R , a packing of f (D) is provided as f (L ∩ D). N (?) means that the metric on f (D) ⊂ R induced from the Euclidean metric of N R is diagonal. In a local sense, diagonalization of the metric is possible for any C∞ Riemannian manifolds of dimensions n ≤ 3. The case of n = 2 follows from the existence of conformal metric λ(x,y)(dx2 + dy2), and the case of n = 3 was proved in [15]. The main purpose of Sections 3 and 4 is to discuss and answer the following: Problem 2 Determine the system of partial differential equations (PDEs) that pro- vide f satisfying (?), (??) for n = N, and make new kinds of aperiodic pack- ings from solutions of the PDEs (Section 3).

Problem 3 Explain the self-similarity observed among the solutions as a general property of the system of PDEs. (Section 4). A family of the PDE solutions can be obtained from solutions of the inviscid Burgers equation ut + uux = 0 (Example 4). Some elementary PDE solutions are also provided in Section 3.2. The result includes packings of a plane with logarith- mic spirals and a packing of a ball as a 3D analogue of the Vogel spiral. In Section 4, after mentioning the general case of n ≤ N (i.e., packing of local charts of the Riemannian manifolds), first we prove that any real analytic Riemann surface has an atlas {(Uα ,ϕα )}α∈I such that every Uα has the metric our pack- ing method can be applied to (Theorem 3). The same thing is also true for any piecewise differentiable , because (?) and (??) for n = 1 means that f is a parametrization of f (D) by the arc-length. Theorem 4 is our attempt to deduce the self-similarity observed among the PDE solutions. In particular, all the presented packings are the images of lattice points contained in a rectangle D (or a rectangular parallelepiped for the 3D case), which suggests that there is an inductive way to construct them. Another commonly ob- served property is that if the last entry xn is separated so that x = (xn−1,xn) and n−1 D = D2 × I (D2 ⊂ R , I ⊂ R), every f can be represented by

α(h(x)) f (x) = e U(h(x)) f2(xn−1) + v0 (3)

N for some functions h : D → R, α : h(D) → R, U : h(D) → O(N), f2 : D2 → R N and v0 ∈ R . Thus, if t = h(x) is considered as the time variable, the image of {x ∈ D : t = h(x)} by f has the identical shape f2(D2) for any t ∈ I. In order to explain the self-similarity, we shall provide a method to construct (n+1)-dimensional Riemannian manifolds with a diagonal and constant-determinant

4 metric from manifolds of dimension n with such a metric (Theorem 4). Theorem 4 can provide a large family of the PDE solutions for dimensions n = 2,3. As one of the examples, packings of shell-surfaces represented by the Raup model in shell morphology [36], are presented in Example 8. The Vogel spiral and the phyllotaxis model ((b) of Figure 2) generated by the golden angle method, can be regarded as a model of a growing disk and a growing cylindrical surface, respectively. Our discrete packing method can also provide a simple mathematical model of growing cylinders and balls, accretive growth of shell-surfaces, and thus, all the types of biological growth particularly mentioned in [19]. It is interesting to ask to what extent the solutions of the PDE system of Theorem 2 generally works as a local model of biological growth. Aperiodic packing has a number of applications. In addition to mesh gener- ation and biological pattern formation mentioned above, when quasicrystals were discovered [42], the Penrose tiling [35] was immediately suggested as a mathe- matical model of aperiodic structures with a long-range order [27]. More recently, iterative algorithms for generating circle packings on Riemann surfaces have been studied [11], inspired by the Koebe-Andreev-Thurston theorem [24], [4], [45] and the Thurston conjecture proved in [39]. We believe that our packing method which has its origins in botany, will broaden the range of applications of geometry of numbers.

Methods to color point sets

All the presented packings are 2D or 3D scatter plots displayed by Wolfram Math- ematica. The points are colored by either of two methods. The first one colors each point according to the local packing density in its neighborhood (e.g., Figure 3). For aperiodic packing, there is ambiguity in defining the local density. However, in the considered situation in which the lattice basis B and the map f are specified, the local density around f (x) can be approximated as the density of the lattice packing given by the lattice L(J(x)B), where J(x) is the Jacobian matrix of the map f (x). The second one colors each point, according to its birth time, by considering some xi among x = (x1,...,xn) as the time when the point f (x) is generated (e.g., Figure 8). This method allows easy observation of the self-similarity hidden in the image.

Notation and symbols

The symbol ∝ signifies that the left-hand and right-hand sides are equal up to a constant multiple. The composite function f (g(x)) is denoted by ( f ◦ g)(x). O(n) n is the orthogonal group of degree n. For any u,v ∈ R , the Euclidean inner product is represented by u · v.

5 The continued fractions are represented using squared brackets [] as follows.

1 1 1 1 [a0,a1,a2,··· ,an,···] := a0 + ··· ··· . a1+ a2+ a3+ an+

If [a0,a1,...] is purely periodic, there is an integer m > 0 such that am+n = an for any n ≥ 0. In such a case, [a0,a1,...] is abbreviated as [a0,a1,...,am−1]. The n’th convergent pn/qn of ϕ = [a0,a1,a2,···] is the ratio of the coprime integers pn, qn that satisfy pn/qn = [a0,a1,a2,··· ,an]. n For any lattice L ⊂ R of full rank, we denote the packing density by ∆(L).

πn/2 (min L)n/2 ∆(L) = , (4) Γ(n/2 + 1) 2nvol(Rn/L) n where Γ(x) is the gamma function, Vol(R /L) is the volume of the fundamental n domain R /L, and min L is the squared length of the shortest vectors.

min L := {|l|2 : 0 6= l ∈ L}. (5)

B ∈ GLn(R) is called a basis matrix of L, if the columns of B are a basis of L. Such a lattice L is also denoted as L(B) explicitly. For any diagonal D ∈ GLn(R), the lattice L(DB) is also denoted by D · L. We define the Lagrange number and Markoff spectrum for Theorem 1.

Definition 1. For any α, the supremum supM of M that fulfills (*) is called the Lagrange number of α, and denoted by L (α). 1 (*) Infinitely many rationals p/q satisfy |α − p/q| < . Mq2

The set {L (α) : α ∈ R \ Q} is called the Lagrange spectrum. For any irrational α, its Lagrange number can be calculated from the contin- ued fraction expansion α = [a0,a1,a2,...] by the following formula (cf. Proposi- tion 1.22, [3]):

L (α) = limsup([an+1,an+2,...] + [0,an,an−1,...,a1]). n→∞

For any indefinite binary quadratic form f (x,y) = ax2 + bxy + cy2 with real coefficients, its discriminant d( f ) and minimum m( f ) are defined by:

d( f ) = b2 − 4ac,  2 m( f ) = min | f (x,y)| : 0 6= (x,y) ∈ Z . p Definition 2. The set of M ( f ) := d( f )/m( f ) of all the indefinite binary quadratic forms over R is the Markoff spectrum.

6 If α is a quadratic irrational (i.e., Q(α) is a quadratic field over Q), the conju- gate of α is denoted by α. The Markoff theorem states that the Lagrange spectrum and the Markoff spectrum coincide below 3 [29]. In fact, any indefinite binary quadratic forms f over R with M ( f ) < 3, has a root α of f (x,1) = 0 that fulfills M ( f ) = L (α). Such an α (i.e., α ∈ R with L (α) < 3) is a quadratic irrational with the continued fraction expansion α = [a0,...,an,γ], where γ is equal to one of the following γm: √ m + 2u + 9m2 − 4 γ = , (6) m 2m where m is one of the Markoff numbers m = 1,2,5,13,... i.e., positive integers that fulfill the Markoff equation for some integers m1,m2: 2 2 2 m + m1 + m2 = 3mm1m2. The integer 0 ≤ u ≤ m/2 is the solution of u2 ≡ −1 mod m.

2 Parastichies from a viewpoint of lattice-basis reduction and Markoff theory

Mathematically, parastichies are the images of lines that connect lattice points with the shortest vectors. In order to see this more precisely, it is useful to review Markov theory from a viewpoint of the lattice theory. Proposition 1 used in the proof of Theorem 1, describes how the Selling re- duced basis, including the shortest vectors, varies when the scales of the x and y axes are changed. Although similar results have been known in the study of indef- inite quadratic forms, Proposition 1 cannot be obtained directly from them. Let ϕ1 > 0 > ϕ2 be real numbers with the continued fraction expansions:

ϕ1 = [a0,a1,a2,··· ,an,···], −1 −ϕ2 = [a−1,a−2,··· ,a−n,···]. ∞ The doubly infinite sequence {an}n=−∞ associated with (ϕ1,ϕ2) is obtained as a result. For simplicity, ϕ1,ϕ2 are assumed to be irrational, and hence an > 0 for any n 6= 0,−1. (+) (+) (−) (−) Let pn /qn = [a0,a1,a2,··· ,an], pn /qn = [a−1,a−2,a−3,··· ,a−n−1] be −1 the n’th convergent of ϕ1 and −ϕ2 , respectively. For n = −1,−2, we put: ! ! p(+) p(+) p(−) p(−) 1 0 −1 −2 = −1 −2 = . (+) (+) (−) (−) 0 1 q−1 q−2 q−1 q−2

If it is assumed that [a0,a1,...] is purely periodic, then ϕ1 = [a0,a1,...,am−1] is quadratic over R, and the following equality holds for the conjugate ϕ1 of ϕ1 (Lemma 1.28, [3]): −1/ϕ1 = [am−1,am−2,...a0].

7 ∞ Therefore, if ϕ2 = ϕ1, {an}n=−∞ is also a periodic sequence.

For the above ϕ1,ϕ2 and any ε > 0, let Lϕ1,ϕ2,ε be the lattice generated by the following basis:

 −1/2  −1/2  ε −ε ϕ1 b0 := 1/2 , b−1 := 1/2 . ε −ε ϕ2

A new basis bn,bn−1 of Lϕ1,ϕ2,ε is provided by: ( p(+) b + q(+) b if n > 0, b = n−1 0 n−1 −1 n n (−) (−) (−1) (q−n−2b0 − p−n−2b−1) if n < −1. In fact, from the definition,  !   p(+) p(+)  n−1 n−2  b0 b−1 (+) (+) if n ≥ 0,   qn−1 qn−2 bn bn−1 = !   q(−) −q(−)  n −n−2 −n−1 (−1) b0 b−1 (−) (−) if n ≤ 0,  −p−n−2 p−n−1  ! ! ! ! ε−1/2 0 1 −ϕ a 1 a 1  1 0 ··· n−1 if n ≥ 0,  1/2  0 ε 1 −ϕ2 1 0 1 0 = ! ! ! ! ε−1/2 0 1 −ϕ 0 1 0 1  1 ··· if n ≤ 0.  1/2  0 ε 1 −ϕ2 1 −a−1 1 −an

Hence,  !   a 1  n−1  bn−1 bn−2 if n ≥ 1,   1 0 bn bn−1 = !   0 1   bn+1 bn if n ≤ −1.  1 −an

Thus, bn = an−1bn−1 + bn−2 holds for any n ∈ Z. From a0,a−1 ≥ 0, an > 0 ∞ (n 6= 0,−1), the following implies that {bn · bn−1}n=−∞ increases monotonically:

bn · bn−1 − bn−1 · bn−2 = an−1bn−1 · bn−1.

In addition,

 −1 (+) (+)  (+) (+)  (+) (+)  ε qn−1qn−2 pn−1/qn−1 − ϕ1 pn−2/qn−2 − ϕ1      (+) (+) (+) (+) (+) (+)  +εqn−1qn−2 pn−1/qn−1 − ϕ2 pn−2/qn−2 − ϕ2 n ≥ 0, bn ·bn−1 = 2 (−) (−)  (−) (−) −1 (−) (−) −1 −εϕ2 q−n−1q−n−2 p−n−1/q−n−1 + ϕ2 p−n−2/q−n−2 + ϕ2      −1 2 (−) (−) (−) (−) −1 (−) (−) −1  −ε ϕ1 q−n−1q−n−2 p−n−1/q−n−1 + ϕ1 p−n−2/q−n−2 + ϕ1 n ≤ 0.

8 (+) The first terms converge to 0 as n → ∞. The second terms diverge owing to qn → (−) ∞ and q−n → ∞ (n → ∞). Hence, lim bn · bn−1 = ∞, and lim bn · bn−1 = −∞ are n→∞ n→−∞ obtained. The term superbasis was first used in [12] to explain the Selling reduction [41].

Definition 3. For any lattice L of rank 2, a basis u1,u2 ∈ L and u3 = −u1 − u2 are a Selling-reduced superbasis, if u1,u2,u3 satisfy u1 · u2 ≤ 0, u1 · u3 ≤ 0, and u2 · u3 ≤ 0.

Proposition 1. For any ϕ1 > 0 > ϕ2 and ε > 0, suppose that a lattice Lϕ1,ϕ2,ε and its basis vectors bn (n ∈ Z) are defined as above. Let N be the integer that fulfills n ≥ N ⇔ bn ·bn−1 ≥ 0, and d be the smallest integer that satisfies (dbN−1 +bN−2)· bN−1 ≥ 0. In this case, the following u1,u2,u3 are a Selling reduced superbasis of

Lϕ1,ϕ2,ε .

u1 := bN−1, u2 := (d − 1)bN−1 + bN−2, u3 := −u1 − u2 = −(dbN−1 + bN−2).

In addition, one of the following holds:

(i) |u1| < |u2|,|u3|.

(ii) |u2| ≤ |u1| < |u3|. In this case, d = 1, u2 = bN−2.

(iii) |u3| ≤ |u1| < |u2|. In this case, d = aN−1, u3 = −bN.

(iv) |u2|,|u3| ≤ |u1|. In this case, d = aN−1 = 1 and u2 = bN−2, u3 = −bN.

In particular, one of bN−2,bN−1, or bN is the shortest vector of Lϕ1,ϕ2,ε .

Proof. From the choice of N, bN · bN−1 ≥ 0 and bN−1 · bN−2 < 0 hold. Since (xbN−1 +bN−2)·bN−1 monotonically increases as a function of x, there is an integer 1 ≤ d ≤ aN−1 as stated above. We shall show that u1,u2,u3 are a Selling-reduced superbasis; clearly, u1,u2 are a basis of Lϕ1,ϕ2,ε . From the definition, u1 · u2 < 0 and u1 · u3 ≤ 0. Hence, we only need to show that u2 ·u3 ≤ 0. For the proof, N ≥ 1 may be assumed; in fact, if ϕ1,ϕ2 are replaced by [a−M,a−M+1,...,], −[0,a−M−1,a−M−2,...] for some −M < N, and   ε is replaced accordingly, then b0 b−1 is replaced by b−M b−M−1 , and N is set to a positive integer. From the process of the continued fraction expansion, aN (N ≥ 1) is the largest integer among all the c’s for which the following two have different signatures:

p(+) N−1 − = [a ,a ,...,a ] − , (+) ϕ1 0 1 N−1 ϕ1 qN−1 cp(+) + p(+) N−1 N−2 − = [a ,a ,...,a ,c] − . (+) (+) ϕ1 0 1 N−1 ϕ1 cqN−1 + qN−2

9 In particular, the following two have the same signature for any 1 ≤ r ≤ aN: (r − 1)p(+) + p(+) N−1 N−2 − = [a ,a ,...,a ,r − ] − , (+) (+) ϕ1 0 1 N−1 1 ϕ1 (r − 1)qN−1 + qN−2 rp(+) + p(+) N−1 N−2 − = [a ,a ,...,a ,r] − . (+) (+) ϕ1 0 1 N−1 ϕ1 rqN−1 + qN−2

Thus, u2 · u3 ≤ 0 is obtained from the following: (+) (+) (+) (+) −u2 · u3 = ((d − 1)qN−1 + qN−2)(dqN−1 + qN−2)   (+) (+)  (+) (+)   (+) (+)  (+) (+)  −1 (d−1)pN−1+pN−2 dpN−1+pN−2 (d−1)pN−1+pN−2 dpN−1+pN−2 × ε (+) (+) − ϕ1 (+) (+) − ϕ1 + ε (+) (+) − ϕ2 (+) (+) − ϕ2 . (d−1)qN−1+qN−2 dqN−1+qN−2 (d−1)qN−1+qN−2 dqN−1+qN−2

Therefore, u1,u2,u3 is Selling reduced, which implies that one of u1,u2,u3 is the shortest vector of L. As for the second statement, the following equation is obtained in a similar way as the above: (+) (+) ! (+) (+) ! (d − 2)p + p dp + p |u | ≤ |u | ⇔ −(u − u ) · u ≤ 0 ⇒ N−1 N−2 − N−1 N−2 − < 0, (7) 2 1 2 1 3 (+) (+) ϕ1 (+) (+) ϕ1 (d − 2)qN−1 + qN−2 dqN−1 + qN−2 (+) (+) ! (+) (+) ! (d + 1)p + p (d − 1)p + p |u | ≤ |u | ⇔ −u · (u − u ) ≤ 0 ⇒ N−1 N−2 − N−1 N−2 − < 0. (8) 3 1 2 3 1 (+) (+) ϕ1 (+) (+) ϕ1 (d + 1)qN−1 + qN−2 (d − 1)qN−1 + qN−2 Because the values in the parentheses of Eq.(7) have different signatures owing to the inequality, |u2| ≤ |u1| implies d = 1. Similarly, |u3| ≤ |u1| implies d = aN−1.

0 0 0 0 The determination of ∆n, ∆ (L(Bn)) is equivalent to that of the λn,λ (Bn). 0 0 λn := sup λ (Bn), Bn∈GLn(R),detBn=±1 0 λ (Bn) := inf min L(DBn). D∈GLn(R): diagonal detD=±1

In fact, if the determinant of Bn ∈ GLn(R) is fixed to ±1, the following holds from Eq.(4): (πλ 0(B )/4)n/2 ∆0(L(B )) = n . (9) n Γ(n/2 + 1) 0 0 Determining λn and λ (Bn) is reduced to the problem about λn and λ(Bn), which is known as products of linear forms.

λn := sup λ(Bn), (10) Bn=(bi j)∈GLn(R), detB=±1

n λ(Bn) := inf (bi1x1 + ··· + binxn) . n ∏ 06=(x1,...,xn)∈Z i=1

10 The main part of Theorem 1 can be reduced to the following lemma.

Lemma 1. (1) For any B = (bi j) ∈ GLn(R) with detB = ±1 and λ(Bn) 6= 0, the following equality holds:

0 2/n λ (Bn) = n(λ(Bn)) . (11)

(2) If a totally real field K of degree n has discriminant dK, −1/2 0 then, λn ≥ dK . As a result, for each n, some Bn attains the supremum ∆n. Furthermore,

(πλ 0/4)n/2 (πn/4)n/2λ (πn/4)n/2d−1/2 ∆0 = n = n ≥ K . n Γ(n/2 + 1) Γ(n/2 + 1) Γ(n/2 + 1)

Proof. (1) From the inequality of arithmetic and geometric means, the part ≥ is proved since we have the following:

n 0 2 2 λ (Bn) = inf ∑ di (bi1x1 + ··· + binxn) , d1,...,dn∈R, d1···dn=±1, n i=1 06=(x1,...,xn)∈Z

n λ(Bn) = inf (bi1x1 + ··· + binxn) . n ∏ 06=(x1,...,xn)∈Z i=1

n Furthermore, > is impossible, because if |∏i=1(bi1x1 + ··· + binxn)| = v for n some 0 6= (x1,...,xn) ∈ Z , the left-hand side of Eq.(11) cannot be more than 2/n 1/n nv , which is seen by putting di = v /(bi1x1 + ··· + binxn) .

(2) Let σ1,...,σn be distinct embeddings of K into C over Q, b1,...,bn be a basis of the ring oK of integers of Kn over Z,√ and Bn ∈ GLn(R) be the matrix with σi b j in the (i, j)-entry. From |detBn| = dK and min σ1(α)···σn(α) = 1, 06=α∈oK the following is obtained:

−1/2n −1/2 λn ≥ λ(dK Bn) = dK .

This implies λn > 0, namely, the following star body is finite type.

n {(x1,...,xn) ∈ R : |x1 ···xn| ≤ 1}.

Hence, λ(Bn) = λn holds for some Bn ∈ GLn(R) with detBn = ±1 (Theorem 0 0 0 0 9 of §17, chap.3 in [20]). For such a Bn, λ (Bn) = λn and ∆ (L(Bn)) = ∆n hold owing to Eqs.(9) and (11). The last inequality is also clear.

For comparison, the packing density of the densest lattice packings are: √ • n = 2: π/2 3 ≈ 0.907,

11 √ • n = 3: π/3 2 ≈ 0.740 [17],

• n = 4: π2/16 ≈ 0.617 [25], √ • n = 5: π2/15 2 ≈ 0.465 [25].

In each dimension, the Voronoi algorithm can be used to find the densest lattice packings [47].

σi Theorem 1. (1) For n = 2–5, let Bn ∈ GLn(R) be the matrix with b j in each (i, j)-entry, where σ1,...,σn are all the embeddings of the following Kn into C over Q, and b1,...,bn are a basis of the ring of integers of the fields Kn as a Z-module. √ −1 2π −1/5 • K2 = Q(ζ5 + ζ5 ), ζ5 = e , √ −1 2π −1/7 • K3 = Q(ζ7 + ζ7 ), ζ7 = e , p √ • K4 = Q( 7 + 2 5), √ −1 2π −1/11 • K5 = Q(ζ11 + ζ11 ), ζ11 = e ,

For the above Bn, the following hold: √ 0 0 π 0 0 3π ∆ = ∆ (L(B2)) = √ ≈ 0.702, ∆ = ∆ (L(B3)) = ≈ 0.389, 2 2 5 3 14 √ 2 2 0 0 π 0 0 5 5π ∆ ≥ ∆ (L(B4)) = √ ≈ 0.183, ∆ ≥ ∆ (L(B5)) = ≈ 0.076. 4 10 29 5 12 · 112

(2) For any distinct ϕ1,ϕ2 ∈ R \ Q, let Lϕ1,ϕ2 ,Lϕ1,ϕ2,ε be the lattices L(B1),L(Bε ) generated by the column vectors of the following matrices:

   −1/2   1 −ϕ1 ε 0 1 −ϕ1 B1 = , Bε = 1/2 . 1 −ϕ2 0 ε 1 −ϕ2

Then,  π liminf∆ Lϕ1,ϕ2,ε = , (12) ε→+0 2L (ϕ1) 0  π ∆ (Lϕ ,ϕ ) = inf∆ Lϕ ,ϕ ,ε = , (13) 1 2 ε 1 2 2M ( f ) p where L (ϕ1) is the Lagrange number of ϕ1, and M ( f ) := d( f )/m( f ) is the element of the Markoff spectrum that corresponds to the quadratic form f (x,y) := (x − ϕ1y)(x − ϕ2y).

Remark 1. Eq.(12) does not hold for any rational ϕ1, because L (ϕ1) = 0 if ϕ1 ∈ Q (Corollary 1.2, [3]).

12 Remark 2. The relation between parastichies and Markoff theory seems to have been first explicitly described in [8]. In fact, their Theorem 1 can be considered to handle a special case of Eq.(12), as their glowth capacity is a constant multiple of the packing density of the lattice packing with the basis in Eq.(1). √ Remark 3. The upper bounds λ4 ≤√3/(20 5) [49] and λ5√≤ 1/57.02 [18] are 0 2 0 2 also known. Therefore, ∆4 ≤ 3π /(40 5) ≈ 0.331 and ∆5 ≤ 5 5π /(12·57.02) ≈ 0.161.

Proof. √(1) This part is an immediate consequence√ of the known results: λ2 = −2 1/ 5 [25], [28], λ3 = 1/7 [14], λ4 ≥ 1/5 29 [31] and λ5 ≥ 11 [22]. The lower bounds for λ4,λ5 are from the smallest discriminants among all the to- tally real quartic and quintic fields.

(2) As a result of (2) of Lemma 1, Eq.(13) is obtained as follows:

π inf f (x,y) n πλ(B ) 06=(x ,...,xn)∈ πm( f ) π 0(L ) = 1 = 1 Z = = . ∆ ϕ1,ϕ2 p 2|detB1| 2|ϕ1 − ϕ2| 2 d( f ) 2M ( f )

Eq.(12) is proved as follows; ϕ1 > 0 > ϕ2 may be assumed, by replacing Bε with DBε g for some for some diagonal D ∈ GL2(R) and g ∈ GL2(Z). This replaces each ϕi by a linear fractional transformation of ϕi. Hence, it does not change their Lagrange numbers.

From Proposition 1, for sufficiently small ε > 0, the shortest vector of Lϕ1,ϕ2,ε (+) (+) (+) (+) is equal to pN b1 −qN b2 using the N’th convergent pN /qN of ϕ1. Further- more, the following is proved as in Lemma 1.

−1 (+) (+) 2 (+) (+) 2 liminfminLϕ1,ϕ2,ε = liminf inf {ε (pN − qN ϕ1) + ε(pN − qN ϕ2) } ε→+0 ε→+0 N>0

(+) (+) (+) (+) = 2liminf (pN − qN ϕ1)(pN − qN ϕ2) N→∞

(+) (+) (+) = 2|ϕ1 − ϕ2|liminf qN (pN − qN ϕ1) N→∞ 2|ϕ − ϕ | = 1 2 , (cf. proof of Proposition1.22 of [3]) limsupN→∞ `N(ϕ1) `n(ϕ1) := [an+1,an+2,...] + [0,an,an−1,...,a1].

Thus, Eq.(12) is proved as follows:

 π π liminf∆ Lϕ1,ϕ2,ε = = . ε→+0 2limsupN→∞ `N(ϕ1) 2L (ϕ)

As an immediate consequence of Theorem 1, it is possible to improve the pack- ing around the origin of the Vogel spiral by using the optimal lattice basis B2 in

13 0.80 packing density 0.75

0.70

0.65

0.60

Figure 3: Original Vogel spiral (left) and a packing obtained from the lattice basis B2 in Theorem 1 (right).

Theorem 1 (Figure 3). More obvious examples will be provided in Figure 7 of Section 3.2. The real numbers ϕ with L (ϕ) < 3 have been enumerated by Markoff theory.

Example 1 (Vogel spirals for quadratic irrationals ϕ with L (ϕ) < 3). The γm p 2 defined by Eq.(6) has the Lagrange number L (γm) = 9 − 4/m . Hence, the   1 −γm basis matrix Bm = fulfills the following for any Markoff number m. 1 −γm

0 π π π ∆ (L(Bm)) = = p > ≈ 0.5236. 2L (γm) 2 9 − 4/m2 6

√ The quadratic irrationals√ with the smallest√ Lagrange numbers are γ1 = (1 + 5)/2 = [1], γ2 = 1 + 2 = [2], γ5 = (9 + 221)/10 = [2,2,1,1]. For each m = 1,2,5, the packing density of L(DBm) is not less than the following value, regardless of the diagonal D ∈ GL2(R):

0 π ∆ (L(B1)) = √ ≈ 0.7025, 2 5 0 π ∆ (L(B2)) = √ ≈ 0.5554, 4 2 0 5π ∆ (L(B5)) = √ ≈ 0.5283. 2 221

Figure 4 presents the Vogel spirals obtained as the image of L(B j) ( j = 2,5). The map f will be provided in Example 5.

3 Application to packings of the Euclidean spaces

3 n In this section, it is assumed that n = N, and f (x) = ( f1(x),..., fn(x)) ∈ C (D,R ) n defined on open subset D ⊂ R , satisfies the properties (?), (??):

14 0.8 packing density

0.7

0.6

0.5

Figure 4: Vogel spirals obtained as the images of L(B2) (left) and L(B5) (right). The formulas used for all the figures are summarized in Table 2 of Appendix B.

(?) For any x ∈ D, there are an orthogonal matrix U(x) of degree n and a diago- nal matrix Φ(x) with the diagonal entries φi(x) > 0 (i = 1,...,n) that satisfy the following:   ∂ f1/∂x1 ··· ∂ f1/∂xn  . .  J(x) :=  . .  = U(x)Φ(x). ∂ fN/∂x1 ··· ∂ fN/∂xn

(??) detΦ(x) = c for some constant c > 0.

After the system of the PDEs for the above Φ(x) and U(x) is determined in Theorem 2, a family of the PDE solutions are provided by using solutions of invis- cid Burgers equation (Section 3.1) and separation of variables (Section 3.2).

3.1 System of PDEs and solutions provided by inviscid Burgers equa- tion First, only (?) is assumed. The constraint (??) is not used until Example 3. Since 3 n t 2 f (x) ∈C (D,R ) and the Jacobian matrix J(x) fulfills JJ = Φ(x) , φ1(x),...,φn(x) 2 −1 are C functions. Let Ak(x) be the matrix defined by Ak(x) := U(x) (U(x))xk . t From ( UU)xk = O and (Uxk )x j = (Ux j )xk , the following are obtained: t (a) Ak + Ak = O.

(b) A jAk − AkA j = (A j)xk − (Ak)x j . The entries of Φ(x) can be specified as a solution of the following second-order nonlinear PDEs.

15 n Theorem 2. Let D ⊂ R be a simply-connected open subset. The φ1(x),...,φn(x) n that fulfill (?) for some f ∈ C3(D,R ), are provided by solving the following PDEs:

−1 −1 φi (φi)xk (φ j)xi + φk (φk)xi (φ j)xk = (φ j)xixk (1 ≤ i, j,k ≤ n : distinct), (14) −1   −1  −2 φ (φ j)x + φ (φk)x = − φ (φk)x (φ j)x (1 ≤ j,k ≤ n : distinct). (15) k k xk j j ∑ i i i x j i6= j,k

 (k) For fixed Φ(x), the matrix Ak = ai j (k = 1,...,n) that satisfy the above (a) and (b) are provided by:

t t Ak = ekck − ckek, (16)

−1 −1  where ck := φ1 (φk)x1 ,...,φn (φk)xn and ek is the unit vector with 1 in the k’th entry. U(x) ∈ O(n) is obtained by solving Uxk = UAk (k = 1,...,n).

Proof. If such an f exists, from (( fi)x j )xk = (( fi)xk )x j , φ1(x),...,φn(x) and U(x) =  u1(x) ··· un(x) satisfy:

(k) ( j) (φ j)xk u j + φ j ∑ ai j ui = (φk)x j uk + φk ∑ aik ui (1 ≤ j,k ≤ n). i6= j i6=k  Since the columns of u1(x),...,un(x) ∈ O(n) are linearly independent for any x ∈ D, the following are obtained:

(k) ( j) φ jai j = φkaik (1 ≤ i, j,k ≤ n : distinct), (k) −1 ak j = φ j (φk)x j (1 ≤ j,k ≤ n : distinct).

(k) (k) (k) It is concluded that ai j = 0 for any distinct 1 ≤ i, j,k ≤ n, owing to ai j = −a ji and (k) φk ( j) φk ( j) φk (i) φk (i) (k) ai j = aik = − aki = − ak j = a jk = a ji . φ j φ j φi φi t t Thus, Ak = ekck − ckek is obtained. Because Ak also fulfills the above (b),

t t t t t t t t ( e jc j − c je j)( ekck − ckek) − ( ekck − ckek)( e jc j − c je j) −1 t t t t −1 t t = φk (φ j)xk ( e jck − cke j) − (c j · ck)( e jek − eke j) − φ j (φk)x j ( ekc j − c jek) t t t t = ( e jc j − c je j)xk − ( ekck − ckek)x j . (17) By comparing the ( j,i)- and ( j,k)-components of Eq.(17), we can obtain:

−1 −1  (φiφk) (φ j)x (φk)x = φ (φ j)x (1 ≤ i, j,k ≤ n : distinct), k i i i xk −2 −2 φ j (φ j)x j (φk)x j + φk (φ j)xk (φk)xk − c j · ck −1   −1  = φ (φ j)x + φ (φk)x (1 ≤ j,k ≤ n : distinct). k k xk j j x j Each equation leads to Eqs.(14), (15), respectively.

16 For any constants t1,...,tn 6= 0, f (x1,...,xn) fulfills (?) if and only if f (t1x1,...,tnxn) does. Accordingly, φ j(x1,...,xn) ( j = 1,...,n) fulfill Eqs.(14),(15), if and only if t jφ j(t1x1,...,tnxn)( j = 1,...,n) do. Thus, the self-similar solutions are provided by those that fulfill φ j(x1,...,xn) = t jφ j(t1x1,...,tnxn) for any 0 6= t1,...,tn ∈ R. However, the following f corresponds to the self-similar solutions φ j = c j/x j (0 6= c j ∈ R) of Eqs.(14), (15): does not satisfy (??):   c1 logx1  .  N f (x1,...,xn) = U0  .  + v0 (U0 ∈ O(n),v0 ∈ R ). cn logxn

Next, we discuss the case in which φ1(x) = ··· = φn(x) i.e., f is a conformal map. As in the case of self-similar solutions, the condition (??) only provides trivial lattice packings, because φ1 = ··· = φn implies that Φ(x) and U(x) are constant, as a result of Theorem 2.

Example 2 (Case of conformal mapping). We can put φ(x) := φ1(x) = ··· = φn(x). For n ≥ 3, every φ j(x) must be constant, because any conformal maps for such dimensions are homotheties or congruence transformations. If n = 2, φ fulfills the following equality: (logφ)xx + (logφ)yy = 0. Namely, u(x,y) := logφ(x,y) is a harmonic function. If v(x,y) is harmonic conju- gate to u(x,y) (i.e., ux = vy and uy = −vx), the following is obtained from Eq.(16).

 −1    0 φ φy 0 uy A1 = −1 = , −φ φy 0 −uy 0  −1    0 −φ φx 0 −ux A2 = −1 = , φ φx 0 ux 0 0 −v cosv −sinv U(x,y) = U exp = U , U ∈ O(2). 0 v 0 0 sinv cosv 0

The above example includes the case of the Doyle spirals. In the Doyle spiral, each circle is tangent to six other circles, and the seven circles have the radii with the ratio 1,a,b,1/a,1/b,a/b,b/a as in Figure 5. As seen in Figure 6, the Doyle spirals are the image of a hexagonal lattice in the complex plane C by exponential maps [7]. In [9], conformally symmetric circle packings were defined as a generalization of the Doyle spirals. In order to generate such circle packings, the determination of the radius of each circle is also necessary, which is not the scope of this study. From the side of the golden angle method, packings with the Voronoi cells of varying sizes have been investigated in [43], [48]. The Doyle spirals in Figures 1 and 6 use logarithm circles |logz − α| = r (i.e., circle images by an exponential map), instead of exact circles. Such a numerical construction of the Doyle spiral is also found in [7]. “Apparently hexagonal” circle

17 ーl1 1/A

ーl2 ーl1+l2 0 B/A l ーl 1/B 1 2 l2 1

l1 A/B A B

Figure 5: Left: adjacent circles in hexagonal packing, Right: the ratio of the radii of adjacent circles in the Doyle spiral

Figure 6: Left: hexagonal lattice packing before being mapped, Right: Doyle spiral of type (12,12) and image of the lattice points by a conformal map. packings can be constructed in a similar manner, by mapping an exact hexagonal circle packing with any conformal maps. −1 Example 3 (PDE for n = 2). From Theorem 2, the PDE for n = 2 is (φ2 (φ1)y)y + −1 −1 2 (φ1 (φ2)x)x = 0. If (??) is also assumed, ε(x,y) := c φ2 and the Jacobian matrix U(x,y)Φ(x,y) of f (x,y) : D → R2 must fulfill the following: −1 −1 2 2 εxx + (ε )yy = c ((φ2 )xx + (φ1 )yy) = 0,  −1   −1  0 φ2 (φ1)y 0 (ε )y/2 A1 = −1 = −1 , −φ2 (φ1)y 0 −(ε )y/2 0  −1    0 −φ1 (φ2)x 0 −εx/2 A2 = −1 = . φ1 (φ2)x 0 εx/2 0

From (A1)y − (A2)x = A1A2 − A2A1 = O, some θ(x,y) satisfies εx = 2θy and −1 (ε )y = −2θx.U(x,y) is explicitly represented by using this θ as follows: 0 −θ cosθ −sinθ U(x,y) = U exp = U (U ∈ O(2)). 0 θ 0 0 sinθ cosθ 0 −1 All the solutions of εεx +εy = 0 fulfill εxx +(ε )yy = 0, which is also seen from the following decomposition:  ∂ ∂  ∂ ∂  ε + (ε−1) = − ε−1 + ε−1 ε. xx yy ∂x ∂y ∂x ∂y

18 Non-trivial solutions of Eqs.(14),(15) for dimensions n > 2 are obtained by set- 2 2 ting φ2k−1(x) = ck/εk(x2k−1,x2k), φ2k(x) = ckεk(x2k−1,x2k), (and φn(x) = c(n+1)/2 if n is odd) for some constants ck and solutions of εk(εk)x + (εk)y = 0 (1 ≤ k ≤ (n + 1)/2). −1 Example 4 (Case of inviscid Burgers equation εxx + (ε )yy = 0). Let ε(x,y) be the solution of the following inviscid Burgers equation: ( εε + ε = 0, x y (18) ε(t,0) = h(t) for any t ∈ I, where h(x) : I → R is the initial condition given on an interval I ⊂ R. The map f (x,y) with the following Jacobian matrix, can be determined as fol- lows:     −1/2  ( f1)x ( f1)y cosθ(x,y) −sinθ(x,y) ε 0 = 1/2 . (19) ( f2)x ( f2)y sinθ(x,y) cosθ(x,y) 0 ε

From εεx + εy = 0, −1 −1 2θx = −(ε )y = −ε εx, −1 2θy = εx = −ε εy. Thus, θ = −(logε + d)/2 for some constant d. The inviscid Burgers equation can be solved by using the characteristic equation: ( q0(s) = ε(q(s),s), (20) q(0) = t. From Eq.(20), q0(0) = ε(t,0) = h(t). On the characteristic , q0(s) = ε(q(s),s) is constant because d ε(q(s),s) = q0(s)ε (q(s),s)+ε (q(s),s) = ε(q(s),s)ε (q(s),s)+ε (q(s),s) = 0. ds x y x y Hence, q(s) = t + h(t)s (s ∈ R) is the charasteristic line, on which f = ( f1, f2) satisfies: d  logε + d logε + d  f (q(s),s) = q0(s)( f ) + ( f ) = ε1/2 cos + sin , ds 1 1 x1 1 x2 2 2 d  logε + d logε + d  f (q(s),s) = q0(s)( f ) + ( f ) = ε1/2 cos − sin . dt 2 2 x1 2 x2 2 2 Without loss of generality, d = −π/2 may be assumed. In this case, for any (x,y) ∈ 2 R and t ∈ I that satisfies t + h(t)y = x, f = ( f1, f2) is given by logh(t) ! p sin 2 f (x,y) = f (t,0) + y 2h(t) logh(t) , cos 2  logh(t)−π/2  R √1 cos dt R ( f ) (t,0)dt 2 f (t, ) = 1 x1 = h(t) . 0 R  R 1 logh(t)−π/2  ( f2)x1 (t,0)dt − √ sin dt h(t) 2

19 logh(t) logh(t) In particular, if f1(t,0)cos 2 6= f2(t,0)sin 2 , ! logh(t) logh(t)  Y 1 sin logh(t) f (x,y) = ( f (t,0)cos − f (t,0)sin ) 2 , 1 2 2 2 −1 Y logh(t) cos 2 yp2h(t) + f (t,0)sin logh(t) + f (t,0)cos logh(t) Y = 1 2 2 2 . logh(t) logh(t) f1(t,0)cos 2 − f2(t,0)sin 2

logh(t) logh(t) Otherwise, by using g(t) := f1(t,0)/sin 2 = f2(t,0)/cos 2 ,

logh(t) ! sin 2 p f (x,y) = Y logh(t) , Y = y 2h(t) + g(t). cos 2 Both correspond to the case of Eq.(3), if f is regarded as the function of (t,Y).

3.2 A family of solutions obtained by separation of variables

−1 Another family of solutions of εxx + (ε )yy = 0 can be obtained by separation of variables. If we put ε(x,y) = F(x)/G(y), F00(x)/G(y) = −G00(y)/F(x) is obtained −1 from εxx = −(ε )yy. Hence, for some constant α,

F(x)F00(x) = α, G(y)G00(y) = −α.

0 2 0 2 If α 6= 0, (F (x)) = 2α logF(x) + d1, (G (y)) = −2α logG(y) + d2. In this case, F(x) and G(y) are functions represented by incomplete gamma functions. We will discuss only the case of α = 0 to obtain elementary solutions. If α = 0, F(x) = c1x + d1 and G(y) = c2y + d2 are obtained. By translating the x and y-coordinates, it may be assumed that ε(x,y) = F(x)/G(y) is equal to either of (a) ε = 1/βy, (b) ε = βx, (c) ε = βx/y. In case of (a)–(c), ε satisfies εεx +εy = 0 if and only if (c) and β = 1. Without loss of generality, β > 0 may be assumed. The following examples explains each case. The case (b) is omitted, because it can be obtained from (a) by exchanging x and y, and f1 and f2. The case (a) includes the Vogel spiral. Example 5 ((a) ε(x,y) = 1/βy: packing of a disk). The map f is as follows, up to congruence transformations.

√ cos(βx/2) f (x,y) ∝ y , sin(βx/2) which is injective on the following D:

 2 D := (x,y) ∈ R : 0 ≤ x < 4π/β, 0 < y < M . Hence, the mapped lattice L needs to contain a vector close to t(4π/β,0) so that the point sets near the half-lines x = 0 and x = 4π/β can be smoothly con- nected.

20 As for the basis matrix B of L, we will consider the following:

1 −ϕ 1 −ϕ 0 −1 (i) , (ii) , (iii) , 0 −1 1 −ϕ 1 −ϕ √ In this example, ϕ is always set to 1/(1+γ1) = (3− 5)/2. The above (ii) is same as the optimal B2 of Theorem 1.

(i) If s := 4π/β is an integer, then t(s,0) is precisely contained in L(B). If s = 1, f (L(B) ∩ D) is the same as the Vogel spiral. √ f (L(B) ∩ D) := n(cos(2πnϕ/s),sin(2πnϕ/s)) : n ∈ Z, 0 < n < M .

As shown in (i) of Figure 7, the points f (L(B) ∩ D) are rather sparse around the origin for large s, because ε = s/4πy is not sufficiently small (cf. (2) of Theorem 1). The sparsity can be avoided by using the basis (ii) instead.

(ii) In this case, L(B) does not contain any vectors of the form t(s,0). However, (−) (−) the n’th convergent pn /qn of −1/ϕ fulfills:   (−) ! (−) (−) ! 1 −ϕ qn qn + pn ϕ (−) = (−) (−) 1 −ϕ −pn qn + pn ϕ (−) (−) ! (−) !  2q + p (ϕ + ϕ) (−) pn 1 = n n + q ϕ −ϕ−1 − . n (−) −1 0 qn

(−) (−) Therefore, by setting s to one of 2qn + pn (ϕ + ϕ) (n ≥ 0), it is possible to connect the point sets near the boundary of D apparently smoothly, as seen in (ii) of Figure 7. This technique of setting s to a special value is also used in the other figures.

(iii) As in case (ii), although L(B) does not contain any vectors t(s,0), the bound-

(−) ary problem can be avoided, by putting s = pn , because

  (−) ! (−) ! (−)! 0 −1 qn pn pn (−) = (−) (−) ≈ . 1 −ϕ −pn qn + pn ϕ 0 The packing becomes sparse at the coordinates farther from the origin, as seen in (iii) of Figure 7. The number of spines is equal to the chosen param- eter s.

It is known that the parastichies in the Vogel spiral are the Fermat spiral. In fact, √ the image of (x,y) by the following f has the polar coordinate (r,θ) = ( y,βx/2).

√ cos(βx/2) f (x,y) ∝ y . sin(βx/2)

21 (i) s =47 (ii) s =47

(iii) s =55 (ii) s =45

Figure 7: Packings in a disk obtained from the lattice bases (i)–(iii) of Example 5. (−) (−) (−) The parameter s is set to (i), (ii) s = 2q9 +(γ1 +γ1)p9 = 47, where p9 = −21, (−) √ (−) q9 = 55 are the ninth convergent of −1/ϕ = (−3+ 5)/2, and (iii) s = −p11 = (−) (−) 55. As for (ii), the case of s = 45 6= 2qn +(γ1 +γ1)pn (n ∈ Z>0) is also presented as the case in which the point sets are not smoothly connected around the x > 0 part of the x-axis.

22 √ If x,y is colinear, (r,θ) = ( y,βx/2) satisfies a linear equation r2 = a + bθ for some constants a and b, which is an equation of the Fermat spiral. In the case of (c), the image of any line y = ax passing through the origin by the map f in Eq.(21), is a , because logr and θ fulfill a linear equation. Example 6 ((c) ε(x,y) = βx/y, case of logarithmic spirals). In this case, it is seen 2 that for some U0 ∈ O(2) and v0 ∈ R , √ cosθ(x,y) β −1 β f (x,y) ∝ xyU + v , θ(x,y) = − log|x| + log|y|. (21) 0 sinθ(x,y) 0 2 2 The map f is injective on the D:   2 4π D := (x,y) ∈ R : 0 < logx ≤ , 0 < y ≤ M . β + β −1

It is necessary to set s := exp(4π/(β + β −1)) to a positive integer, as in the previous example. Since X2 − (4π/logs)X + 1 = 0 has a real root β, 1 ≤ s < 2π (−) (−) e ≈ 535.5 is required. In Figure 8, s are set to some s = 2qn + (γ1 + γ1)pn .

Birth time Old New s =3 s =11 s =47

s = 199 s = 521 s =e2π ≈535.5

Figure 8: Packing of planes with logarithmic spirals. Each point is colored accord- ing to the y value of its preimage. If y is regarded as the time variable, the points with y < ymax form the identical shape, regardless of the ymax value. This self- similarity explains their biological shapes (2D snails, 2D embryos, leaves, etc.). The last s = e2π is also the case of an inviscid Burgers solution.

Proposition 2 provides a case of the dimension n = 3.

23 Proposition 2. In the case of n = 3 and A3 = 0, the solutions φ1(x),φ2(x),φ3(x) for the system of PDEs in Eqs.(14), (15) are as follows; for some constants d,d2 −1 2 and ε(x1,x2) with εx1x1 + (ε )x2x2 = −2d2 :

1/3 3 1/3 −1/2 φ1(x) = c (d + 3d2x3) ε (x1,x2), 1/3 3 1/3 1/2 φ2(x) = c (d + 3d2x3) ε (x1,x2), 1/3 3 −2/3 φ3(x) = c (d + 3d2x3) .

−1 Proof. From A3 = U Ux3 = O, U is independent of x3. (φ3)x1 = (φ3)x2 = 0 also −1  −1  holds. From Eq.(15), φ (φ1)x = φ (φ2)x = 0. Thus, if we put G(x3) := 3 3 x3 3 3 x3 R x3 0 φ3(x1,x2,x)dx, the following holds for some Fi(x1,x2) and Hi(x1,x2).

φi = Fi(x1,x2)G(x3) + Hi(x1,x2)(i = 1,2).

From φ1φ2φ3 = c, (φ1φ2)x j = 0 holds for both j = 1,2. Therefore, φ1 = (α1G(x3)+ β1)F˜(x1,x2) and φ2 = (α2G(x3) + β2)/F˜(x1,x2) for some α1,α2,β1,β2 ∈ R and F˜(x1,x2). From Eq.(14) and Eq.(15),

−1 (φ1)x2x3 = φ2 (φ2)x3 (φ1)x2 , (22) −1 (φ2)x1x3 = φ1 (φ1)x3 (φ2)x1 , (23) −1  −1  −2 φ (φ2)x + φ (φ1)x = −φ (φ1)x (φ2)x . (24) 1 1 x1 2 2 x2 3 3 3 From the first two equalities, (log(φ ) ) = (logφ ) , (log(φ ) ) = (logφ ) . 1 x2 x3 2 x3 2 x1 x3 1 x3 Thus, the following functions are independent of x3. G x G x −1 1 α1 ( 3) + β1 ˜ 2 −1 1 α2 ( 3) + β2 ˜ −2 φ2 (φ1)x2 = (F )x2 , φ1 (φ2)x1 = (F )x1 . 2 α2G(x3) + β2 2 α1G(x3) + β1

Hence, it is possible to choose r1,r2 6= 0,d,d2 ∈ R and F˜(x,y) so that φ1 = r1(d + d2G(x3))F˜(x1,x2) and φ2 = r2(d + d2G(x3))/F˜(x1,x2). Thus, the follow- 2 0 0 ing are obtained from φ1φ2φ3 = r1r2(d + d2G(x3)) G (x3) = c, φ3(x3) = G (x3), G(0) = 0 and Eq.(24):

3 3 r1r2(d + d2G(x3)) = r1r2d + 3cd2x3, r r 2 ˜ −2 1 ˜ 2 2 (F )x1x1 + (F )x2x2 = −r1r2d2 . 2r1 2r2 ˜ −2 −1 2 If we put ε(x1,x2) := (r2/r1)F(x1,x2) , εx1x1 +(ε )x2x2 = −2(r1r2)d2 is ob- tained, in addition to the following.

1/3 1/2 3 1/3 −1/2 φ1 = c (r1r2) (d + 3(cd2/r1r2)x3) ε (x1,x2), 1/3 1/2 3 1/3 1/2 φ2 = c (r1r2) (d + 3(cd2/r1r2)x3) ε (x1,x2), 1/3 −1 3 −2/3 φ3 = c (r1r2) (d + 3(cd2/r1r2)x3) .

−1/2 −1/2 The statement is proved if d,d2 are replaced by (r1r2) d, (r1r2) d2/c.

24 Separation of variables can also be used to obtain a family of solutions of −1 2 εx1x1 + (ε )x2x2 = −2d2 for d2 6= 0. A packing of a ball (i.e., 3D analogue of the Vogel spiral), is obtained as a result; if we put ε(x1,x2) = F(x1)/G(x2), εx1x1 + −1 2 (ε )x2x2 + 2d2 = 0 implies:

00 00 2 F(x1)F (x1) + G(x2)G (x2) + 2d2 F(x1)G(x2) = 0.

Hence, F(x1) or G(x2) must be a constant function, and either of (a) ε = 2 2 d0 + d1x1 − (d2x1) or (b) ε = 1/{d0 + d1x2 − (d2x2) } holds. The part (b) can be obtained by swapping the roles of x1 and x2, and f1 and f2 in (a). Therefore, we will discuss only the case (a).

2 Example 7 (Packing of a ball, 3D Vogel spiral). From ε = d0 + d1x1 − (d2x1) , t t d2 6= 0 and Ak = ekck − ckek,

 −1/2   0 0 d2ε 0 −εx1 /2 0 1/2 A1 =  0 0 0 , A2 = εx1 /2 0 d2ε , A3 = O. −1/2 1/2 −d2ε 0 0 0 −d2ε 0

∗ The following D j,Vj ( j = 1,2) provide diagonalizations A1 = V1D1V1 ,A2 = ∗ V2D2V2 of A1 and A2.     i 0 0 −i i √0 −1/2 1 D1 = d2ε 0 −i 0, V1 = √  0 0 2, 0 0 0 2 1 1 0   q i 0 0 2 2 D2 = d1 /4 + d0d2 0 −i 0, 0 0 0 √  1/2 εx /2 εx /2 − 2d2ε 1 q 1 q 1  2 2 2 2  V2 = q −i d1 /4 + d0d2 i d1 /4 + d0d2 0 . 2 2  √  2(d1 /4 + d0d2 ) 1/2 1/2 d2ε d2ε 2εx1 /2

From (Vj)x j = O and Ux j = UA j ( j = 1,2), U(x) satisfies (UVj)x j = UVjD j. Hence, for some U0 ∈ O(3),

1 0 0  1/2  d2ε 0 −εx /2 q q q 1 0 cos d2/4 + d d2x sin d2/4 + d d2x  2 2 U(x) ∝ U0  1 0 2 2 1 0 2 2  0 d /4 + d0d 0 .  q q  1 2  2 2 2 2 1/2 0 −sin d1 /4 + d0d2 x2 cos d1 /4 + d0d2 x2 εx1 /2 0 d2ε

The Jacobian matrix of the map f is provided as U(x)Φ(x). Hence, for some

25 3 v0 ∈ R ,   −εx /2d2 q 1 1/3 3 1/3  1/2 2 2  f (x) ∝ c (d + 3d2x3) U0 ε sin d1 /4 + d0d2 x2  + v0  q  1/2 2 2 ε cos d1 /4 + d0d2 x2  2  d x1 − d1/2 q 2 q 3 1/3  2 2 2 2 2 2  ∝ (d + 3d2x3) U0  d1 /4 + d0d2 − (d2 x1 − d1/2) sin d1 /4 + d0d2 x2  + v0. q q  2 2 2 2 2 2 d1 /4 + d0d2 − (d2 x1 − d1/2) cos d1 /4 + d0d2 x2 q 2 2 By putting s := 1/ d1 /4 + d0d2 ,U0 = I and v0 = 0, and replacing x1, x2, x3 2 3 by (x2/rs + d1/2)/d2 , 2πx1, (x3 − d )/3d2 respectively, the following is obtained.   x2 q 1/3  2 2  f (x) ∝ x  r − x2 sin(2πx1/s), r,s > 0 : constants. 3 q  2 2 r − x2 cos(2πx1/s)

The above f is injective on the D, and maps D onto a ball:

3 D := {(x1,x2,x3) ∈ R : 0 ≤ x1 < s, −r < x2 < r, 0 < x3 < R}.

As the lattice basis, B3 of Theorem 1 can be used (see Table 2 of Appendix A). Figures 9 and 10 present the surface pattern and the cross-sections of the 3D Vogel spiral for the parameters s = 1 and r = R = 1000.

0.60

0.55 packing density

0.50

0.45 z

y 0.40

0.35

Figure 9: Point distribution around the origin (left) and the pattern on the semi- sphere (right) of the 3D Vogel spiral for the parameters s = 1, r = R = 1000.

26 0.60

0.55 packing density

0.50

0.45

0.40 y z 0.35 x y

Figure 10: Cross-sections of the 3D Vogel spiral in the x-y and y-z planes in case of s = 1 and r = R = 1000.

4 General case: packing of the Riemannian manifolds

N Let N ≥ n > 0 be integers, and f (x) = ( f1(x),..., fn(x)) : D → R be a function m defined on an open subset D ⊂ R that fulfills (?), (??). Hence,   ∂ f1/∂x1 ··· ∂ f1/∂xn Φ(x)  . .  = U(x) , O : (N − n) × n zero matrix.  . .  O ∂ fN/∂x1 ··· ∂ fN/∂xn As in the previous section, if n and N are fixed, it is possible to derive the system of PDEs for φ1(x),...,φn(x) and U(x). However, it is also important to determine which Riemannian manifolds can be (locally) packed by the proposed method. Theorem 3 deals with this problem in the real analytic surface case (class Cω ). Although the same thing does not seem to hold for general dimensions, The- orem 4 ensures that the method is applicable to three-dimensional manifolds with the self-similar properity as stated. If the dimension N is appropriately chosen, any Ck Riemannian manifolds N (M,g) can be isometrically embedded into the Euclidean space R by an injective k map of class C (3 ≤ k ≤ ∞ or ω) [32], [33]. We fix an atlas {(Uα ,ϕα )}α∈I of M and N assume that each open subsets Uα ⊂ M has the isometry ια : Uα → ια (Uα ) ⊂ R . If the metric g of (M,g) is locally diagonalizable, it may be also assumed that for n 2 2 any α ∈ I, g|Uα is diagonal ∑i=1 φi(x) dxi so that the Jacobian matrix Jα (x) of −1 ια ◦ ϕα fulfills (?). For another fixed diffeomorphism ψα : Uα → ψα (Uα ), it is straightforward to −1 see that f := ια ◦ψα fulfills (?) and (??) on some open subset V ⊂ ψα (Uα ) if and only if the following ([) holds:

−1 ([) The diffeomorphism x = σ(y) := ϕα ◦ ψα on V has the Jacobian matrix t t Jσ such that Jσ (y) Jα (σ(y))Jα (σ(y))Jσ (y) is diagonal, and has a constant

27 determinant.

t In particular, if ϕα is an isothermal coordinate system, i.e., Jα (x)Jα (x) = t λ(x)I for some positive function λ(x), then ([) occurs if and only if Jσ (y)Jσ (y) is −1 diagonal, and the Jacobian matrix Jσ −1 of the inverse function σ satisfies t −2 2 det( Jσ −1 (x)Jσ −1 (x)) = c λ (x). In Theorem 3, we assume that n = 2, and the Riemann surface (M,g) is real analytic in order to use the Cauchy-Kovalevskaya theorem. Theorem 3. For any constant c 6= 0 and real analytic Riemannian surface (M,g), there is an atlas {(Uα ,ϕα )}α∈I of M and an isometry ια : Uα → RN on each of Uα such that for any α ∈ I, we have a real analytic function ε(x1,x2) on ϕα (Uα ) with

2 c 2 2 g|Uα = dx1 + ε(x1,x2)dx2. ε(x1,x2)

−1 Namely, ια ◦ ϕα fulfills (?), (??).

Proof. For any p ∈ M, fix a neighborhood p ∈ Uα ⊂ M and a diffeomorphism 2 ϕα : Uα → ϕα (Uα ) ⊂ R . We may assume that Uα has an isometry ια : Uα → 2 ια (Uα ) ⊂ R , and ϕα is an isothermal coordinate system. We shall prove that −1 some neighborhood V ⊂ Uα of p and diffeomorphism σ with the inverse σ : 2 −1 ϕα (V) → R (hence, ψα := σ ◦ ϕα |V ) satisfy ([). If so, the chart (V,ψα ) has the −1 desired property, i.e., ια ◦ ψα fulfills (?) and (??). t t −1 −1 −1 t Jσ −1 (x) Jσ −1 (x) = ( Jσ (σ (x))Jσ (σ (x))) is diagonal. and det(Jσ −1 (x) Jσ −1 (x)) = −2 2 c λ (x), if and only if Jσ −1 is represented as follows for some real analytic func- tions ε(x1,x2) and θ(x1,x2),

 1/2  −1 ε (x1,x2) 0 Jσ −1 (x1,x2) = c λ(x1,x2) −1/2 U(x1,x2), 0 ±ε (x1,x2)   cosθ(x1,x2) sinθ(x1,x2) U(x1,x2) = . −sinθ(x1,x2) cosθ(x1,x2)

−1 −1 If we put Bi := (Jσ −1 ) (Jσ −1 )xi and Ai := U Uxi , λ ε 1 0  B = xi I + xi tU U + A i λ 2ε 0 −1 i     λxi εxi cos2θ(x1,x2) sin2θ(x1,x2) 0 1 = I + + θxi . λ 2ε sin2θ(x1,x2) −cos2θ(x1,x2) −1 0

From the symmetry of second derivatives, the second column of Jσ −1 B1 and the first column of Jσ −1 B2 must be equal. Hence, the second column of B1 and the first column of B2 are also equal:             λx1 0 εx1 sin2θ(x1,x2) 1 λx2 1 εx2 cos2θ(x1,x2) 0 + +θx1 = + +θx2 . λ 1 2ε −cos2θ(x1,x2) 0 λ 0 2ε sin2θ(x1,x2) −1

28 Thus, cos2θ(x ,x ) −sin2θ(x ,x )(logε)  −(logλ) + θ  1 2 1 2 x2 = 2 x2 x1 . sin2θ(x1,x2) cos2θ(x1,x2) (logε)x1 (logλ)x1 + θx2

From ((logε)x2 )x1 = ((logε)x1 )x2 ,      −(logλ)x1x2 + 2θx1x1 −(logλ)x2 + 2θx1 1 tan2θ(x1,x2) + 2θx2 (25) (logλ)x1x1 + 2θx1x2 (logλ)x1 + 2θx2      −(logλ)x2x2 + 2θx1x2 −(logλ)x2 + 2θx1 = −tan2θ(x1,x2) 1 − 2θx1 . (logλ)x1x2 + 2θx2x2 (logλ)x1 + 2θx2 From the Cauchy-Kowalevskaya theorem, Eq.(25) has a real analytic solution θ defined on ϕα (V) for some simply-connected neighborhood p ∈ V ⊂ U. From −1 this θ, logε and σ (x1,x2) can be constructed by solving the above PDEs. Next, we prove Theorem 4. For any diffeomorphism σ : U → V between open subsets of two manifolds and metric g on V, For any p ∈ U and u,v ∈ TpU, the pullback metric σ ∗g is defined by ∗ (σ g)p(u,v) = gσ(p)(dσp(u),dσp(v)). Theorem 4. For any integers 0 < n < N, and k = 2,3,...,∞ or ω, let f (x) ∈ k N n C (D,R ) be a function on a simply-connected open subset D ⊂ R with the Ja- cobian matrix Jf (x). Suppose that f (x) satisfies (i) and (ii) for some constant α ∈ R and antisymmetric matrix A of degree N: t ∗ (i) Jf (x)Q(x)Jf (x) = O for any x ∈ D, if we put f (x) := (αI + A) f (x), and: Q(x) := f ∗(x) t f ∗(x)A + A f ∗(x) t f ∗(x) − ( f ∗(x) · f ∗(x))A.

t ∗ (ii) det( Jf (x)Jf (x)) 6= 0, and f (x) is linearly independent of fx1 (x),..., fxn (x) over R for any x ∈ D. k Let Hf (x) ∈ C (D,R) be the function determined by the following equations, except for the constant term: ∗ f (x) · fx (x) (H ) (x) = j ( j = 1,...,n). f x j f ∗(x) · f ∗(x)

Let q f = (qi j)1≤i, j≤n be the positive-definite symmetric matrix.  ∗ t ∗  − 2(n+1) αH (x) ∗ ∗ 1 t f (x) f (x) q (x) := e n f ( f (x) · f (x)) n J (x) I − J (x). f f t f ∗(x) f ∗(x) f n Hence, g f = ∑i, j=1 qi jdxidx j is a Riemannian metric on D. For any function N h(x,xn+1) on D × R>0,Ff ,h(x,xn+1) : D × R>0 → R is defined by

αh(x,xn+1) Ff ,h(x,xn+1) := e exp(Ah(x,xn+1)) f (x). The following (a) and (b) are equivalent:

29 k (a)F f ,h(x,xn+1) satisfies (?), (??) for some h(x,xn+1) ∈ C (D × R>0,R).

(b) The metric g f is diagonal and has a constant determinant on D.

For any Ck diffeomorphism x = σ(y) on D, (a0), (b0) are also equivalent:

0 k (a )F f ◦σ,h(y,xn+1) fulfills (?), (??) for some h(y,xn+1) ∈ C (D × R>0,R). 0 ∗ (b ) The pull-back σ g f = g f ◦σ is diagonal and has a constant determinant on D.

Remark 4. In (b0), such a σ exists whenever n = 1, or n = 2 and k = ω as a result of Theorem 3. In (i), Q(x) = O holds whenever A = O, or n = 1, or N = 2,3 and α = 0.

Proof. From (i),

∗ ∗ ! t ∗ t ∗ ∗ t ∗ ∗ ∗  f · fx   f · f  2 J ( f f A + A f f − ( f · f )A)J j − xi = f f = O. f ∗ · f ∗ f ∗ · f ∗ ( f ∗ · f ∗)2 xi x j i, j=1,...,n

The existence of Hf is obtained from this and the generalized Stokes theorem for 1 t −1 t 1-forms of class C (Theorem 6.1, XXIII, [26]). Since I − Jf ( J f Jf ) Jf is the projection onto the linear space generated by fx1 ,..., fxn , the assumption (ii) im- plies:

−2(n+1)αHf (x) t ∗ t −1 t  ∗ t detq f (x) = e ( f (x) I − Jf ( Jf Jf ) Jf f (x))det( Jf (x)Jf (x)) 6= 0.

Thus, q f (x) is positive-definite, and g f is a Riemannian metric on D. For any diffeomorphism σ on D, (i) and (ii) hold for f if and only if they do for f ◦ σ. Thus, (a0) ⇔ (b0) is immediately obtained from (a) ⇔ (b). Hence, only (a) ⇔ (b) is proved in the following. To show (a) ⇒ (b), we assume that Ff ,h(x,xn+1) satisfies (?) and (??).

αh ∗ (Ff ,h)x j (x,xn+1) = e exp(Ah)(hx j f (x) + fx j (x)), ( j = 1,...,n) αh ∗ (Ff ,h)xn+1 (x,xn+1) = e exp(Ah)hxn+1 f (x).

Because of (??), hxn+1 (x,xn+1) 6= 0 must hold for any (x,xn+1) ∈ D × R>0. In ∗ ∗ addition, (Ff ,h)x j ·(Ff ,h)xn+1 = 0 ( j = 1,...,n) from (?), and f (x)· f (x) 6= 0 from (ii). These imply:

∗ f (x) · fx (x) h (x,x ) = − j . x j n+1 f ∗(x) · f ∗(x)

Therefore, h(x,xn+1) fulfills for some function h0(xn+1):

h(x,xn+1) = −Hf (x) + h0(xn+1). (26)

30 We also have:  f ∗(x) t f ∗(x) (F ) (x,x ) = eαh exp(Ah) I − f (x)( j = 1,...,n). f ,h x j n+1 t f ∗(x) f ∗(x) x j

Thus, using the Jacobian matrix Jf (x) of f (x),

 f ∗(x) t f ∗(x) ((F ) · (F ) ) = e2αh tJ (x) I − J (x) (27) f ,h xi f ,h x j 1≤i, j≤n f t f ∗(x) f ∗(x) f n+1 1 2α(h+ Hf ) ∗ ∗ − = e n ( f (x) · f (x)) n q f (x).

F(x,xn+1) fulfills (??) if and only if the following holds for some constant c 6= 0:

n+1 ((F ) · (F ) ) = h2 e2(n+1)α(h+Hf ) detq (x) ∏ f ,h x j f ,h x j xn+1 f j=1

0 2 2(n+1)αh0(xn+1) 2 = (h0(xn+1)) e detq f (x) = c ,

2 which implies: detq f (x) = γ for some γ 6= 0. Thus, (a) ⇒ (b) is proved. Conversely, the above discussion shows that (b) ⇒ (a) is obtained if there is a 0 (n+1)αh0(xn+1) constant d1 6= 0 such that h0(xn+1)e = d1. Namely,  1 −Hf (x) + log(d1xn+1 + d2) if α 6= 0, h(x,xn+1) = (n + 1)α (28) −Hf (x) + d1xn+1 + d2 if α = 0.

In particular, if d1 = 1 and d2 = 0, the above h(x,xn+1) is defined on D × R>0, which proves (b) ⇒ (a).

As seen from Eq.(27), if N = n + 1 and a diffeomorphism σ on D satisfies (b0) 2 of Theorem 4 for some 3 ≤ k ≤ ∞ or k = ω, the following φ j (x,xn+1) ( j = 1,...,n) 2 2 n 2 and φn+1(x,xn+1) = c /∏ j=1 φ j (x,xn+1) are a solution of the PDEs of Theorem 2.

∗ 2 ! ( f (σ(x)))x · f (σ(x)) φ 2(x,x ) = e2αh(x,xn+1) ( f (σ(x))) · ( f (σ(x))) − j , j n+1 x j x j f ∗(σ(x)) · f ∗(σ(x))

where h(x,xn+1) is the function obtained by replacing Hf in Eq.(28) with Hf ◦σ . The 3D Vogel spiral can be obtained from Theorem 4 and the following param- eters as the case of n = 2, N = 3:

0 0 0  t  1/3 α = 0, A = 0 0 1, f (s,t) = s √ 0 . 0 −1 0 r2 −t2

31 The following Example 8 is also a special case of Theorem 4, for which n = 1, N = 3 and A, f are given by:     0 ∓1 0 φ1(t) A = ±1 0 0, f (t) =  0 . (29) 0 0 0 φ2(t)

Example 8 (Shell surfaces given by the Raup model). The Raup model is a clas- sical model of gastropod-like surfaces [36], which is obtained by expanding and rotating a generating curve C : (φ1(t),0,φ2(t)) (tb ≤ t < te) around the z-axis:    coss ∓sins 0 φ1(t) αs F(s,t) = e ±sins coss 0 0  (s ∈ R, tb ≤ t ≤ te), (30) 0 0 1 φ2(t) where the signature ± determines the chirality of the shell surface. t Herein, it is assumed that f (t) = (φ1(t),0,φ2(t)) is a continuous, and piece- wise differentiable function on the interval [ta,tb) that fulfills (i) and (ii) of The- orem 4 for the A in Eq.(29). The formulas for the ellipse case are presented in Appendix B. From Eq.(28), the coordinate transformation of F(s,t) can be given by:

1 s = −H (x ) + logx , t = σ(x ), f ◦σ 1 2α 2 1 where Hf ◦σ (x1) is calculated by:

Z f ∗(σ(x )) · f 0(σ(x ))σ 0(x ) Z φ (s)φ 0 (s) + φ (s)φ 0 (s) H (x ) = 1 1 1 dx = 1 1 2 2 ds . (31) f ◦σ 1 ∗ ∗ 1 α 2 2 2 2 f (σ(x1)) · f (σ(x1) (1 + α )φ (s) + α φ (s) 1 2 s=σ(x1)

The metric q f (t)dt on the generating curve C is provided by:   −4αHf (t) 0 0 ∗ ∗ ∗ 0 2 q f (t) = e ( f (t) · f (t))( f (t) · f (t)) − f (t) · f (t)

−4αHf (t) 2 0 2 0 2 2 0 0 2 = e φ1(s) ((φ1(s)) + (φ2(s)) ) + α (φ1(s)φ2(s) − φ2(s)φ1(s)) .

0 2 2 From detq f ◦σ (x1) = (σ (x1)) detq f (σ(x1)) = γ , the σ is determined by:

Z x1 Z σ(x1) q 0 q detq f (σ(x))σ (x)dx = detq f (s)ds = γ(x1 − xb), xb σ(xb) where xb is the base point of the integral that can be chosen arbitrarily from −1 σ ([tb,te)). Thus, G(σ(x1)) = γ(x1 − xb) holds, if we put: Z t q −2αHf (t) 2 0 2 0 2 2 0 0 2 G(t) := e φ1(s) ((φ1(s)) + (φ2(s)) ) + α (φ1(s)φ2(s) − φ2(s)φ1(s)) ds. (32) σ(xb)

32 Since G(t) is a monotonically increasing function, it is possible to compute the −1 + inverse function G . The following f (x1,x2) fulfills (?) and (??):

+ f (x1,x2) := F(−Hf ◦σ (x1) + (logx2)/2α,σ(x1))  +  φ1(σ(x1))cosθ (x1,x2) + αθ (x1,x2) + = e ±φ1(σ(x1))sinθ (x1,x2). (33) φ2(σ(x1))

+ For any G(tb)/γ ≤ x1 −xb ≤ G(te)/γ and x2 ∈ R>0, the above θ (x1,x2) and σ(x1) are calculated by: 1 θ +(x ,x ) = −H (x ) + logx , 1 2 f ◦σ 1 2α 2 −1 σ(x1) = G (γ(x1 − xb)).

It should be noted that different packings are obtained by: 1 θ −(x ,x ) = H (x ) + logx , 1 2 f ◦σ 1 2α 2 abs 1 θ (x1,x2) = − Hf ◦σ (x1) + logx2. 2α If θ +, α are replaced by −θ −, −α in Eq.(33), and the chirality is also inverted from A to −A, the following f − is obtained:

 −  φ1(σ(x1))cosθ (x1,x2) − − αθ (x1,x2) − f (x1,x2) = e ±φ1(σ(x1))sinθ (x1,x2), φ2(σ(x1))  abs  φ1(σ(x1))cosθ (x1,x2) abs abs αθ (x1,x2) abs f (x1,x2) = e ±φ1(σ(x1))sinθ (x1,x2). φ2(σ(x1))

The above f − and f abs parametrize the same surface as f +, although their packing patterns are different. The pseudo code to output a packing of shell surfaces is presented in√ Table 3 of Appendix B, which uses ellipses as the generating curve. M = 10000/ 5, k = 1, xb = 0 and the parameters in Table 1 were used to calculate the packings in Figure 11. For the parameter set 4, the patterns of f + and f abs are also presented in Figure 12. The difference between the patterns was the most easily observed in this bivalve case.

Acknowledgments

This project was financially supported by the SENTAN-Q program of MEXT Ini- tiative for Realizing Diversity in the Research Environment, JST-Mirai Program

33 Table 1: Raup parameters [37] used for Figures 11,12 taken from a document of the software Raup’s Coiler [16]. See Table 3 of Appendix B for the usage of the parameters. W 2T D S R Case 1 1.4 6.3 0. 0.99 10 Case 2 2. 1.8 0. 2. 10 Case 3 3. 0. 1.5 1. 1 Case 4 10000. 0.1 0. 1.6 10 Case 5 10000. 0. 0. 1. 10

Birth time 2 4 Old New 1

3 5

Figure 11: Packing of shell surfaces (Raup model [36]) with logarithmic spirals. As in Figure 8, each packing maintains the identical shape at any time. The parameters and formulas used are in Tables 1,3.

34 + abs (1) f (2) f © Hans Hillewaert

Tellina fabula

packing density

0.70 0.75 0.80 0.85

Figure 12: Two distinct packings of the identical shell surface (Case 4 of Table 1, left-handed coiling). The right-hand pattern is not differentiable at some coordinates, although the surface provided by the Raup model is smooth. This discontinuity is frequently observed in real clam patterns (Source of photo: https://en.wikipedia.org/wiki/Fabulina_fabula]/media/ File:Angulus_fabula.jpg)

(JPMJMI18GD) and JSPS KAKENHI (19K03628). The first author participated as an RA, and was also employed by the project of JST CREST (JPMJCR1911). The research by the second author was supported by grants from the Research Grants Council of the Hong Kong SAR, China (HKU 17301317 and 17303618). The authors would like to thank Mr. Y. Azama and Ms. C. Ooisi of Kyushu Uni- versity for their help in coding.

Author contributions

The last author designed the project, and performed the majority of the mathemat- ical research. The first author derived the PDEs and coded programs with the last author. The second author supervised the project as the mentor of the SENTAN-Q program.

References

[1] I. Adler. The role of continued fractions in phyllotaxis. Journal of Algebra, 205:227–243, 1998. [2] I. Adler, D. Barabe, and R. V Jean. A history of the study of phyllotaxis. Annals of Botany, 80:231–244, 1997. [3] M. Aigner. Markov’s Theorem and 100 Years of the Uniqueness Conjecture. Springer, 2013.

35 [4] E. M. Andreev. Convex polyhedra in lobacevski˘ ˘i spaces. Mathematics of the USSR-Sbornik, 10(3):413–440, 1970. [5] Emch Arnold. Mathematics and engineering in nature. Popular Science Monthly, 79:450–458, 1911. [6] P. Atela, C. Gole,´ and S. Hotton. A dynamical system for plant pattern forma- tion: a rigorous analysis. Journal of Nonlinear Science, 12:641–676, 2002. [7] Alan F. Beardon, Tomasz Dubejko, and Kenneth Stephenson. Spiral hexago- nal circle packings in the plane. Geometriae Dedicata, 49:39–70, 1994. [8] Franc¸ois Bergeron and Christophe Reutenauer. Golden ratio and phyllotaxis, a clear mathematical link. Mathematical Biology, 78:1–19, 2019. [9] Alexander I. Bobenko and Tim Hoffmann. Conformally symmetric circle packings: a generalization of Doyle’s spirals. Experimental Mathematics, 10(1):141–150, 2001. [10] Louis Bravais and Auguste Bravais. Essai sur la disposition des feuilles curviseri´ ees.´ Annales des sciences naturelles (Botanique), 7:42–110, 1837. [11] C. R. Collins and K. Stephenson. A circle packing algorithm. Computational Geometry, 25:233–256, 2003. [12] J. H. Conway. The Sensual (Quadratic) Form (Carus Mathematical Mono- graphs 26). The Mathematical Association of America, Washington, DC, 1997. [13] H. S. M. Coxeter. The role of intermediate convergents in tait’s explanation for phyllotaxis. Journal of Algebra, 20:167–175, 1972. [14] H. Davenport. On the product of three homogeneous linear forms (ii). Pro- ceedings of the London Mathematical Society, s2-44(1):412–431, 1938. [15] D. M. DeTurck and D. YANG. Existence of elastic deformations with pre- scribed principal strains and triply orthogonal systems. Duke Mathematical Journal, 51(2):243–260, 1984. [16] Claudio Fanelli and J. Bartolome. Raup’s coiler, 2013. [17] C. F. Gauss. Untersuchungen uber¨ die Eigenschaften der positiven ternaren¨ quadratischen Formen von Ludwig August Seeber, L. A. J. reine angew. Math., 20:312–320, 1840. [18] H. J. Godwin. On the product of five homogeneous linear forms. Journal of the London Mathematical Society, s1-25:331–339, 1950. [19] A. Goriely. The Mathematics and Mechanics of Biological Growth. Springer- Verlag, New York, 1 edition, 2017.

36 [20] P.M. Gruber and C.G. Lekkerkerker. Geometry of Numbers (North-Holland Mathematical Library 37). North-Holland, Amsterdam, 2 edition, 1987.

[21] D. P. Hardin, T. Michaels, and E. B. Saff. A comparison of popular point configurations on S2. Dolomites Research Notes on Approximation, 9:16–49, 2016.

[22] J. Hunter. The minimum discriminant of quintic fields. Proc. Glasgow Math. Assoc., 3:57–67, 1957.

[23] R. V. Jean. Number-theoretic properties of two-dimensional lattices. Journal of Number Theory, 29:2O6–223, 1988.

[24] P. Koebe. Kontaktprobleme der Konformen Abbildung. Ber. Sachs.¨ Akad. Wiss. Leipzig, Math.-Phys. Kl., 88:141–164, 1936.

[25] A. Korkine and G. Zolotareff. Sur les formes quadratiques. Mathematische Annalen, 6:366–389, 1873.

[26] Serge Lang. Real and functional analysis. Springer-Verlag, 3 edition, 1993.

[27] A. L. Mackay. Crystallography and the Penrose pattern. Physica A, 114:609– 613, 1982.

[28] A. Markoff. Sur les formes quadratiques binaires indefinies.´ Mathematische Annalen, 15:381–406, 1879.

[29] A. Markoff. Sur les formes quadratiques binaires indefinies.´ Mathematische Annalen, 17:379–399, 1880.

[30] A. M. Mathai and T. A. Davis. Constructing the sunflower head. Mathemati- cal Biosciences, 20:117–133, 1974.

[31] J. Mayer. Die absolut-kleinsten Diskriminarten der biquadratischen Zahlkorper.¨ S. B. Akad. Wien, Ila, 138:733–742, 1929.

[32] John Nash. The imbedding problem for riemannian manifolds. Annals of Mathematics, 63(1):345–355, 1956.

[33] John Nash. Analyticity of the solutions of implicit function problem with analytic data. Annals of Mathematics, 84(3):20–63, 1966. √ [34] M. Naylor. Golden, 2, and π flowers: a spiral story. Mathematics Magazine, 75(3):163–172, 2002.

[35] R. Penrose. Set of tiles for covering a surface, 1975.

[36] D. M. Raup. Computer as aid in describing form in gastropod shells. Science, 138(3537):150–152, 1962.

37 [37] D. M. Raup. Geometric analysis of shell coiling: general problems. Journal of Paleontology, 40(5):1178–1190, 1966. [38] J. N. Ridley. Ideal phyllotaxis on general surfaces of revolution. Mathemati- cal Biosciences, 79(1):1–24, 1986. [39] B. Rodin and D. Sullivan. The convergence of circle packings to the Riemann mapping. Journal of Differential Geometry, 26(2):349–360, 1987. [40] J.-F. Sadoc, J. Charvolin, and N. Rivier. Phyllotaxis on surfaces of constant gaussian . Journal of Physics A: Mathematical and Theoretical, 46(29):295202, 2013. [41] E. Selling. Ueber die binaren¨ und ternaren¨ quadratischen Formen. J. reine angew. Math., 77:143–229, 1874. [42] D. Shechtman, I. Blech, D. Gratias, and J. W. Cahn. Metallic phase with long- range orientational order and no translational symmetry. Physical Review Letters, 53(20):1951–1953, 1984. [43] T. Sushida and Y. Yamagishi. Geometrical study of phyllotactic patterns by Bernoulli spiral lattices. Development, Growth & Differentiation, 59(5):379– 387, 2017. [44] R. Swinbank and R. J. Purser. Fibonacci grids: A novel approach to global modelling. Quarterly Journal of the Royal Meteorological Society, 132:1769–1793, 2006. [45] W. Thurston. The geometry and topology of 3-manifolds. Princeton lecture notes, 1978–1981. [46] H. Vogel. A better way to construct the sunflower head. Mathematical Bio- sciences, 44:179–189, 1979. [47] G. F. Voronoi. Sur quelques propriet´ es´ des formes quadratiques positives parfaites. J. reine angew. Math., 133:97–178, 1907. [48] Y. Yamagishi and T. Sushida. Spiral disk packings. Physica D, 345:1–10, 2017. [49]G. Zilinskas.˘ On the product of four homogeneous linear forms. Journal of the London Mathematical Society, s1-16(1):27–37, 1941.

A Formulas for the packings in figures

n Herein, the basis matrix B of lattice L and map f : D → R used to obtain each figure are summarized. As for the Jacobian matrix J(x) of f (x), a constant multiple of tJJ is also included in Table 2, because the local packing density around each point is approximated as the density of the lattice packing L(J(x)B).

38 Table 2: Parameters and formulas used for the packings with constant det tJJ √ Figure 4 √(Example 1): Vogel spirals for the Markoff quadratic irrationals γ2 = 1 + 2, γ5 = (9 + 221)/10: 1 −γ  B = m (m = 2,5), 1 −γm f ,D, tJJ : case of s = 1 of Figure 7.

1 −ϕ Figure 7 (Example 5): packings of a disk for the lattice bases (i) B = , (ii) 0 −1 1 −ϕ 0 −1 √ B = , (iii) B = , where ϕ := 1/(1 + γ ) = (3 − 5)/2: 1 −ϕ 1 −ϕ 1

√ cos(2πx/s) f (x,y) = y , s > 0 : integer, sin(2πx/s) 2 D = {(x,y) ∈ R : 0 < x ≤ s, 0 < y < R}, 4πy/s 0  ( tJJ)(x,y) ∝ . 0 s/4πy

Figure 8 (Example 6): packings of a plane made with logarithmic-spiral patterns:

B : same as (ii) of Figure 7, β : root of X2 − (4π/logs)X + 1 = 0, 0 < s < e2π ≈ 535.5 : integer, √ cosθ(x,y) β −1 β f (x,y) = xy , θ(x,y) = − logx + logy, sinθ(x,y) 2 2 2 D = {(x,y) ∈ R : 1 < x ≤ s, 0 < y < R}, y/βx 0  ( tJJ)(x,y) ∝ . 0 βx/y

Figures 9, 10 (Example 7): packing of a ball (3D Vogel spiral):

 −1 −1 2  1 ζ7 + ζ7 (ζ7 + ζ7 ) √ 2 −2 2 −2 2 2π −1/7 B = 1 ζ7 + ζ7 (ζ7 + ζ7 )  (ζ7 = e ), 4 −4 4 −4 2 1 ζ7 + ζ7 (ζ7 + ζ7 )   1 0 0  x2 f (x ,x ,x ) = x1/3 0 cos(2 x /s) sin(2 x /s)  0 , s > 1 : integer, 1 2 3 3  π 1 π 1 q  − ( x /s) ( x /s) 2 2 0 sin 2π 1 cos 2π 1 r − x2 3 D = {(x1,x2,x3) ∈ R : 0 ≤ x1 < s, −r < x2 < r, 0 < x3 < R},  2 2 2 2/3 2  4π r − x2 x3 /s 0 0 ( tJJ)(x ,x ,x ) ∝  2 2/3 2 2 . 1 2 3  0 r x3 /(r − x2) 0  2 4/3 0 0 r /9x3 Figure 11, 12 (Example 8): Raup surface packing:

B : same as (ii) of Figure 7, f ,D, tJJ : case of k = 1 of Table 3.

39 B Pseudo code for the Raup surface packings

We formulate the case in which the generating curve is an ellipse:

(φ1(x1),0,φ2(x1)) = (X0 + cosσ(x1),0,Y0 + Ssinσ(x1)) (−π ≤ x1 ≤ π).

By using some xb ∈ R that fulfills σ(xb) = θ1(xb) = 0, Hf ◦σ in Eq.(31) is given by:

Z σ(x1) 0 0 Z tan(σ(x1)/2) φ1(s)φ1(s) + φ2(s)φ2(s) P(t) Hf ◦σ (x1) = α 2 2 2 2 ds = 2α dt (t = tan(s/2)), σ(xb) (1 + α )φ1(s) + α φ2(s) 0 Q(t) where

2 2 2 2 P(t) = (−2X0t + SY0(1 −t ))(1 +t ) + 2(S − 1)t(1 −t ), 2 2 2 2 2 2 2 2 Q(t) = ((1 + α )(X0(1 +t ) + 1 −t ) + α (Y0(1 +t ) + 2St) )(1 +t ).

Let R(t) be the polynomial that fulfills Q(t) = R(t)R(t)(1 +t2):

p 2 2 p 2 R(t) = 1 + α (X0 + 1) + iαY0 + 2iαSt +t ( 1 + α (X0 − 1) + iαY0). Z s P(t) In this case, F˜(s) := dt can be calculated by: 0 Q(t)  " #s 2  P(τ )  2  P(τ )  i(τ − τ ) 2t − τ − τ  j  s j j j j j ∑ Re 0 log (t − τ j)(t − τ j) 0 + 2 ∑ Im 0 arctan  Q (τ j) Q (τ j) τ j − τ j τ j − τ j  j=1 j=1 0   if X0 6= 1 or Y0 6= 0,  2 2 2 ˜ 1 − S 1 − α (1 − S ) F(s) = − [log1 +t2]s − [log1 + α2 + α2S2t2]s  2(1 + α2(1 − S2)) 0 4α2(1 + α2(1 − S2)) 0   if X = 1,Y = 0 and α2 6= −1/(1 − S2),  0 0 1 − S2 (1 − S2)2  1 s  [ +t2]s − X = ,Y = 2 = − /( − S2),  log 1 0 2 2 if 0 1 0 0 and α 1 1 4 2S 1 +t 0 (34) where τ0 = i, and τ1,τ2 are the roots of R(t) = 0 when X0 6= 1 or Y0 6= 0: q √ 2 2 2 2 −iαS ± 1 + α (1 − S ) − ( α + 1X0 + iαY0) τ1,τ2 = √ . 2 α + 1(X0 − 1) + iαY0

The G(t) values of Eq.(32) are calculated by numerical integration: Z t q 2 2 2 2 2 2  G(t) := (X0 + coss) (1 + (S − 1)cos s) + α (SX0 coss +Y0 sins + S) exp −4α F˜(tan(s/2)) ds. 0

In case of f − and f abs, the above F˜(tan(s/2)) is replaced by −F˜(tan(s/2)) and abs − F˜(tan(s/2)) . As a result, the values of f (x1,x2) can be calculated more nu- merically stably than f + and f −, because the argument of the is always negative.

40 For simplicity, the following vertical strip D including (xb,0) is used as the domain of f +, f − and f abs, although whether the maps are injective on the D, depends on the choice of α,X0,Y0,S.   2 G(−π) G(π) D = (x1,x2) ∈ R : ≤ x1 − xb < , m < x2 < M . (35) γ γ

Because each segment {(x1,c2) : xb +G(−π)/γ ≤ x1 < xb +G(π)/γ} (c2: con- stant) in D is mapped onto a closed curve, the lattice L is required to contain a vec- tor close to t((G(π) − G(−π))/γ,0) for smoothly connecting the packing patterns near the end points. In the field of shell morphology, the parameters α,X0,Y0,m and M are normally determined by using the other parameter set (see Tables 1, 3).

41 Table 3: Algorithm for the Raup surface packing (case of the generating curve given by φ1(t) = X0 + cost, φ2(t) = Y0 + Ssint (−π ≤ t ≤ π)) (Input) W : expansion rate of the generating curve, /* W,T,D are the Raup parameters [37]. */ T : translation rate, D : distance of the generating curve from the z-axis, S : shape parameter of the ellipse, R : number of revolutions, B : 2-by-2 basis matrix, /* The generated lattice is L(B). */ M : positive integer to specify the number of points in the output figure, N : positive integer to divide the interval [−π,π], k : positive integer, /* L is assumed to contain a vector close to (k,0). */ xb ∈ R : base point of integrals.

(Algorithm) logW 1 + D 2T (1) Set α,X ,Y by α = , X = , Y = . 0 0 2π 0 1 − D 0 1 − D (2) Make a list of pairs (G(t j),t j) with t j = −π + 2π j/N ( j = 0,...,N), where G : [−π,π] → R is the increasing function defined by Eq.(32) and θ1(xb) = −1 σ(xb) = 0. (In what follows, the inverse function G : [G(−π),G(π)] → R is calculated by interpolating this list.) (3) Set xmin,xmax,ε0,ymin,ymax by: kG(−π) kG(π) xmin = x + , xmax = x + , b G(π) − G(−π) b G(π) − G(−π) k(1 − e−4παR) ε = , y = ε−1e−4παR, y = ε−1. 0 M|detB| min 0 max 0

In this case, D in Eq.(35) given by D = [xmin,xmax) × (ymin,ymax). (4) For each of (x,y) ∈ L(B) ∩ D, compute the following:

 G(π) − G(−π)  σ(x) = G−1 (x − x ) , k b

θ1(x) = −2αF˜(tan(σ)/2), /* F˜ in Eq.(34) */ 1 θ(x,y) = ±θ (x) + log(y), /* The sign can be specified arbitrarily. */ 1 2α then plot: /* + is the case of right-handed coiling. */

 cosθ(x,y) 0 φ (σ(x)) 0  f (x,y) = eαθ(x,y) ±sinθ(x,y) 0 1 .   0 φ (σ(x)) 0 1 2

At the same (x,y), the Jacobian matrix J fulfills:

 (G(π) − G(−π))2e2αθ(x,y)  0 t  k2{(1 + 2) ( (x))2 + 2 ( (x))2}  JJ =  α φ1 σ α φ2 σ .  e−2αθ(x,y){(1 + α2)φ (σ(x))2 + α2φ (σ(x))2}  0 1 2 4α2 The approximated value of the packing density near f (x,y) can be obtained as the density of the lattice packing L(JB).

42