Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2022, Volume 213, Issue 11, Pages 1530–1558
DOI: https://doi.org/10.4213/sm9727e
(Mi sm9727)
 

Counting lattice triangulations: Fredholm equations in combinatorics

S. Yu. Orevkov

Steklov Mathematical Institute of Russian Academy of Sciences, Moscow, Russia
References:
Abstract: Let $f(m,n)$ be the number of primitive lattice triangulations of an $m\times n$ rectangle. We compute the limits $\lim_n f(m,n)^{1/n}$ for $m=2,3$. For $m=2$ we obtain the exact value of the limit, which is $(611+\sqrt{73})/36$. For $m=3$ we express the limit in terms of a certain Fredholm integral equation for generating functions. This provides a polynomial-time algorithm (with respect to the number of computed digits) for the computation of the limit with any prescribed precision.
Bibliography: 13 titles.
Keywords: asymptotic, primitive triangulation, Fredholm's integral equation.
Funding agency Grant number
Ministry of Science and Higher Education of the Russian Federation 075-15-2022-265
This work was performed at the Steklov International Mathematical Center and supported by the Ministry of Science and Higher Education of the Russian Federation (agreement no. 075-15-2022-265).
Received: 30.01.2022 and 17.07.2022
Russian version:
Matematicheskii Sbornik, 2022, Volume 213, Number 11, Pages 50–78
DOI: https://doi.org/10.4213/sm9727
Bibliographic databases:
Document Type: Article
MSC: Primary 05A15; Secondary 45B05
Language: English
Original paper language: Russian

§ 1. Introduction

A lattice triangulation of a (lattice) polygon in $\mathbb{R}^2$ is a triangulation with all vertices in $\mathbb{Z}^2$. As discovered in [3], lattice triangulations are important in algebraic geometry (also see [9]). A lattice triangulation is called primitive (or unimodular) if each triangle is primitive, that is, has the minimum possible area $1/2$ (is a translate of $[(0,0),(x_1,y_1),(x_2,y_2)]$ with $x_1y_2-x_2y_1=1$). We denote the number of primitive lattice triangulations of an $m\times n$ rectangle by $f(m,n)$. Let

$$ \begin{equation*} c(m,n)=\frac{\log_2f(m,n)}{mn}, \qquad c_m=\sup_n c(m,n)=\lim_{n\to\infty} c(m,n) \end{equation*} \notag $$
and
$$ \begin{equation*} c=\sup_m c_m=\lim_{m\to\infty} c_m =\sup_n c(n,n)=\lim_{n\to\infty} c(n,n). \end{equation*} \notag $$
The existence of the limits was proved in [4], Proposition 3.6. In [4] the number $c(m,n)$ was called the capacity of an $m\times n$ rectangle. In [8] I gave an upper bound $c<6$ (which can easily be improved to $c<\log_2 27= 4.755$ using the same arguments: using the notation of [8], it is enough not to distinguish the cases $v_j=1$ and $v_j=2$). Later on, a much better estimate $c<3$, as well as ${c_m\mkern-2mu<\mkern-2mu3-1/m}$,

was obtained by Anclin [1]. A still better upper bound $c\mkern-2mu<\mkern-2mu4\log_2\frac{1+\sqrt5}2\mkern-2mu=\mkern-2mu\log_2 6.854\mkern-2mu=2.777$ was obtained in [7] and announced in [12]. (I have not seen the manuscript [7] but Professor Welzl has kindly sent me the slides of his talk [13], where the proof of this bound is clearly exposed.)

It is easy to see that

$$ \begin{equation} f(1,n)=\binom{2n}{n} \quad\text{so that } c_1=2, \end{equation} \tag{1} $$
which yields a lower bound $c\geqslant 2$. It was also computed in [4] that $c\geqslant c_4\geqslant c(4,32)=2.055702$. The following is written in [4], § 2.1: “For $f(2,n)$ we have no explicit formula, and we cannot evaluate the asymptotics precisely”. We still have no explicit formula for $f(2,n)$ but we give here the principal term of the asymptotics.

Theorem 1.1. The following equality holds:

$$ \begin{equation*} \lim_{n\to\infty}f(2,n)^{1/(2n)}=\sqrt\alpha=4.148440\dots, \quad\textit{where } \alpha=\frac{611+\sqrt{73}}{36}. \end{equation*} \notag $$
Hence
$$ \begin{equation*} c_2=\frac{1}{2}\log_2\alpha=2.05256897. \end{equation*} \notag $$

An exact value of $c_3$ in a certain sense is given in Proposition 4.5, where we express $c_3$ in terms of Fredholm’s integral equations for certain generating functions. In particular, Proposition 4.5 provides an algorithm to compute $c_3$ to $n$ digits in polynomial time in $n$. A Mathematica code implementing the main step of this algorithm is presented at the end of § 4.

Theorem 1.2. Up to $360$ digits, the limit

$$ \begin{equation*} \lim_{n\to\infty} f(3,n)^{1/(3n)}, \end{equation*} \notag $$
is equal to
$$ \begin{equation*} \begin{aligned} \, 4.&239369481548025671877625742045235772100695711251795499830801 \\ &687833358238276728987837054831763341276708855553395893005289 \\ &580195934799338289257489707990192054275721787374165246347114 \\ &466096241741151814326914780021501337938335813142441896953051 \\ &597942032082556780952912032761797534112146994900056374798271 \\ &988378451540168358202181556482461979420039542105330977266751, \end{aligned} \end{equation*} \notag $$
and therefore $c_3=2.0838497\dots$ .

We have computed $c_3$ with this high precision in hope to find an algebraic equation for it or relate it to some known constants, but we have not succeeded so far.

In § 2.2 we present the results of computations of the exact values of the numbers $f(m,n)$ for some $m$ and $n$. These computations show, in particular, that $c\geqslant c(5,115)= 2.10449551\dots$ .

In § 6 we give an asymptotic upper bound for the number of all (not necessarily primitive) lattice triangulations. However, it seems to be far from optimal.

The rest of this paper (apart from §§ 2.2 and 6) is devoted to the proof of Theorems 1.1 and 1.2. Namely, in § 2.1 recurrence formulae for the number of primitive lattice triangulations of polygons lying in strips of fixed width are presented. They are similar to those in [4] but, in our opinion, are simpler and more efficient. Theorems 1.1 and 1.2 are proved in § 3 and § 4, respectively, using the relations between generating functions coming from the recurrence relations in § 2.1. As mentioned already, $c_3$ is expressed in terms of a Fredholm integral equation for the Cauchy integrals of the generating functions. In § 5 we estimate the approximation error of the most straightforward numerical solution of a Fredholm equation with analytic kernel.

Acknowledgement

I thank the referee for their valuable comments and corrections.

§ 2. Recurrence relations for strips of fixed width

2.1. Recurrence relations

Given a polygon $P\subset\mathbb{R}^2$, the upper part of its boundary is the set $\{(x,y)\in P\mid y'>y\Rightarrow (x,y')\notin P\}$. A vertical side of $P$ is a side of $P$ contained in a line $\{x=x_0\}$. Let $\mathcal T$ be a triangulation of a polygon $P$ in $\mathbb{R}^2$.

We say that $Q$ is a tile of $\mathcal T$ in the following three cases:

A polygon is called $y$-convex if its intersection with any line $x=\mathrm{const}$ is either an empty set, a point, or a line segment.

Lemma 2.1. Let $\mathcal T$ be a triangulation of a $y$-convex polygon $P$ in $\mathbb{R}^2$. Then there exists a tile $Q$ of $\mathcal T$ such that the upper part of the boundary of $Q$ is contained in the upper part of the boundary of $P$.

Proof. Let $\Gamma_P$ be the upper part of the boundary of $P$. Let $Q_1,\dots,Q_n$ be all the tiles of $\mathcal T$ that have at least one side lying on $\Gamma_P$. Let $\Gamma_i$ be the union of the sides of $Q_i$ lying on $\Gamma_P$. It is clear that each $\Gamma_i$ is either a side of $Q_i$ or a union of two sides with a common vertex. It is also clear that the projections of the $\Gamma_i$ onto the $x$-axis have pairwise disjoint interiors, hence we may assume that $\Gamma_1,\dots,\Gamma_n$ are numbered from left to right.

We say that a tile $Q_i$ is shadowed on the left (shadowed on the right) if the upper part of the boundary of $Q_i$ contains a line segment $I$ such that $I\not\subset\Gamma_P$ and $I$ is on the left (on the right, respectively) of $\Gamma_i$; see Figure 1. It is clear that none of the tiles $Q_1,\dots,Q_n$ can be shadowed on the left and right simultaneously. Hence we may assume without loss of generality that at least one of these tiles is not shadowed on the right. Let $i_0$ be the smallest number such that $Q_{i_0}$ is not shadowed on the right. Then $Q_{i_0}$ is the required tile with upper part contained in $\Gamma_P$. Indeed, it is not shadowed on the right by its definition. It cannot be shadowed on the left either because otherwise $Q_{i_0-1}$ would not be shadowed on the right, which contradicts the minimality of $i_0$.

The lemma is proved.

Now we fix an integer $m>0$ and consider primitive lattice triangulations of polygons which lie in the vertical strip $\{0\leqslant x\leqslant m\}$ and are bounded by two graphs of continuous piecewise linear functions.

By analogy with the terminology introduced in [4], § 2.2, we say that $\varphi\colon [0,m]\!\to\!\mathbb{R}$ is an admissible function if it is a continuous piecewise linear function whose graph is a union of line segments with endpoints in $\mathbb{Z}^2$. We fix an admissible function $\varphi_0$ and say that a function $\varphi\colon [0,m]\to\mathbb{R}$ is $\varphi_0$-admissible if it is admissible and $\varphi(x)\geqslant \varphi_0(x)$ for any $x\in[0,m]$. A $\varphi_0$-admissible shape is a polygon $S$ of the form $\{(x,y)\in\mathbb R^2\mid 0\leqslant x\leqslant m,\, \varphi_0(x)\leqslant y\leqslant\varphi(x)\}$ for some $\varphi_0$-admissible function $\varphi$.

As in the above definition of a tile of a triangulation, we say that $Q$ is a primitive lattice tile in the following three cases:

A primitive lattice tile $Q$ is $P$-maximal for a polygon $P$ if $Q\subset P$ and the upper part of the boundary of $Q$ is contained in the upper part of the boundary of $P$. We say that $S'$ is a $\varphi_0$-admissible subshape of a $\varphi_0$-admissible shape $S$ if $S'$ is the closure of $S\setminus(Q_1\cup\cdots\cup Q_n)$, where $Q_1,\dots,Q_n$ are $S$-maximal primitive lattice tiles with pairwise disjoint interiors. Following [4], in this case we set $\#(S',S)=n$. Note that the relation ‘be an admissible subshape’ is not transitive.

Let us denote the number of primitive lattice triangulations of a polygon $P$ by $f^*(P)$. When $P$ lies in the strip $\{0\leqslant x\leqslant m\}$, we also define $f(P)$ as the number of primitive lattice triangulations of $P$ that have no interior edges whose projections onto the $x$-axis are the whole interval $[0,m]$ (we choose simpler notation for a more complicated notion because the numbers $f(P)$ will be used more often than $f^*(P)$).

The following lemma is the inclusion-exclusion formula in our setting. The proof is the same as for Lemma 2.2 in [4].

Lemma 2.2. For any $\varphi_0$-admissible shape $S$,

$$ \begin{equation*} f^*(S)=\sum_{S'} (-1)^{\#(S',S)-1}f^*(S') \quad\textit{and}\quad f(S)=\sum_{S'} (-1)^{\#(S',S)-1}f(S'), \end{equation*} \notag $$
where the left-hand sum is taken over all proper $\varphi_0$-admissible subshapes of $S$, and the right-hand sum is taken over those proper $\varphi_0$-admissible subshapes of $S$ the upper part of whose boundary contains a point in $\mathbb{Z}^2\cap\{0<x<m\}$.

Example 2.3. Let $m=2$ and $\varphi_0=0$. Given nonnegative integers $a$, $b$ and $c$, let $S_{a,b,c}$ be the $\varphi_0$-admissible shape bounded above by the line segments $[(0,a),(1,b)]$ and $[(1,b),(2,c)]$. Let $f_{a,b,c}=f(S_{a,b,c})$. We also set $f_{a,b,c}=0$ when $\min(a,b,c)<0$. Then (see Figure 2) the recurrence formula in Lemma 2.2 reads

$$ \begin{equation*} f_{a,b,c}= \begin{cases} f_{a-1,b,c}+f_{a,b-1,c}+f_{a,b,c-1}-f_{a-1,b,c-1} &\text{if }(a,b,c)\ne(0,0,0), \\ 1 &\text{if }(a,b,c)=(0,0,0). \end{cases} \end{equation*} \notag $$

Let $F(x,y,z)=\sum_{a,b,c}f_{a,b,c}x^ay^bz^c$ be the generating function. Then summing the recurrence relations over all triples $(a,b,c)\ne(0,0,0)$ we obtain

$$ \begin{equation*} \begin{aligned} \, &F(x,y,z)-1 = \sum f_{a-1,b,c}x^ay^bz^c+\sum f_{a,b-1,c}x^ay^bz^c+\dotsb \\ &\quad = \sum f_{a,b,c}x^{a+1}y^bz^c+\sum f_{a,b,c}x^ay^{b+1}z^c+\dotsb= F(x,y,z)(x+y+z-xz) \end{aligned} \end{equation*} \notag $$
so that $F(x,y,z)=1/(1-x-y-z+xz)$.

Example 2.4. Let $\varphi_0$ and $S_{a,b,c}$ be as in Example 2.3. Given nonnegative $a$ and $c$ such that $a\equiv c+1\mod 2$, we define $S'_{a,c}$ as the $\varphi_0$-admissible shape bounded above by the line segment $[(0,a),(2,c)]$. Let $f^*_{a,b,c}=f^*(S_{a,b,c})$ and $g^*(a,c)=f^*(S'_{a,c})$. We also set $f^*_{a,b,c}=0$ for $\min(a,b,c)<0$ and $g^*(a,c)=0$ for $\min(a,c)<0$ and for $a\equiv c \mod2$. Then for $(a,b,c)\ne(0,0,0)$ the recurrence formula in Lemma 2.2, as applied to $S_{a,b,c}$, reads

$$ \begin{equation*} f^*_{a,b,c}= f^*_{a-1,b,c}+f^*_{a,b-1,c}+f^*_{a,b,c-1}-f^*_{a-1,b,c-1} +\chi_{a,b,c}g^*_{a,c}, \end{equation*} \notag $$
where $\chi_{a,b,c}=1$ if $2b=a+c+1$ and $\chi_{a,b,c}=0$ otherwise. Let $F^*(x,y,z)$ and $G^*(x,z)$ be the corresponding generating functions. Then (cf. Example 2.3) we have
$$ \begin{equation*} F^*(x,y,z)-1=F^*(x,y,z)(x+y+z-xz)+\sum\chi_{a,b,c}g^*_{a,c}x^ay^bz^c, \end{equation*} \notag $$
and the sum on the right-hand side is equal to
$$ \begin{equation*} \sum_{a,c} g^*_{a,c} x^a y^{(a+c+1)/2} z^c =y^{1/2}\sum_{a,c} g^*_{a,c} (xy^{1/2})^a (y^{1/2}z)^c =y^{1/2}G^*(xy^{1/2},y^{1/2}z), \end{equation*} \notag $$
which gives us the relation
$$ \begin{equation*} F^*(x,y,z)(1-x-y-z+xz)=1+y^{1/2} G^*(xy^{1/2},y^{1/2}z). \end{equation*} \notag $$
Now let us apply the recurrence relation to $S'_{a,c}$. The only admissible subshape of $S'_{a,c}$ is $S_{a,(a+c-1)/2,c}$, hence the relation for $S'_{a,c}$ reads $g^*_{a,c} = f^*_{a,(a+c-1)/2,c}$. In terms of the generating functions this means that
$$ \begin{equation*} \begin{aligned} \, &G^*(x,z) = \sum_{a,c} f^*_{a,(a+c-1)/2,c}x^a z^c =\operatorname{coef}_{u^0}\biggl[\sum_{a,b,c} f^*_{a,b,c}x^a u^{2b-a-c+1} z^c\biggr] \\ &\qquad=\operatorname{coef}_{u^0}\biggl[u\sum_{a,b,c} f^*_{a,b,c}(xu^{-1})^a (u^2)^b (zu^{-1})^c\biggr] =\operatorname{coef}_{u^0}\bigl(u F^*(xu^{-1},u^2,zu^{-1})\bigr). \end{aligned} \end{equation*} \notag $$

2.2. Some exact values of $f(m,n)$

The recurrence relations in Lemma 2.2 provide an algorithm for the computation of the exact values of $f(m,n)$ for small $m$ and $n$. The algorithm is similar to the one described in [4], § 2.2. We have performed computations using this algorithm, and the reader can see in Table 1 that we have advanced much further in comparison to the computations in [4]. There have been three reasons for this, whose impacts were more or less equal.

Table 1

Capacities computed in [4]Capacities computed in this paper
$c_1=2.0000$$c_1=2.0000$
$c_{2,375}=2.0441$$c_2=2.0526$
$c_{3,60} =2.0275$$c_3=2.0838$
$c_{4,32}=2.0557$$c_{4,200}=2.0946$
$c_{5,12}=2.0175$$c_{5,115}=2.1045$
$c_{6,7} =1.9841$$c_{6,50\;} =2.1024$
$c_{7,20}=2.0813$
$c_{8,13}=2.0669$
$c_{9,9\;} =2.0490$

The first reason (an obvious one) is that computers have become more powerful. The second reason is that we have used another definition of admissible shapes, which enables us to reduce the amount of memory used with coefficient $3^{m-1}$, which is rather important for $m=9$ (as pointed out in [4], for this kind of algorithm ‘the bottleneck in the computations is always memory’). The third reason is that, instead of long arithmetics, we have used computations modulo different prime numbers and then have recovered the results using the Chinese remainder theorem. This trick has enabled us to ‘convert’ memory into time, the lack of which is not so crucial.

We have computed $f(3,n)$ to $n=600$ and $f(4,n)$ to $n=200$. The exact value of $f(3,600)$ has 1127 digits, and it yields $c_{3,600}=2.07966\dots$ . Comparing this with the limit value $c_3=2.08385$ we see that convergence is very slow. For $m=4$ the last exact value computed is

$$ \begin{equation*} \begin{aligned} \, f(4,200)=\, &262199334303965073140522141167072596609151907003573304927487 \\ &419128543906730659218480439253346584137204205604500628092962 \\ &697997426095545403404830271634194339979807927812812142668569 \\ &097560203843935394728621308903256950859658838687531965864231 \\ &570521446370439565640979852878302993978768696718322811686043 \\ &307749541067654061321020767838164602474781629699981105797912 \\ &385346265396601164596410043968216134349971638142523003353406 \\ &530183843913302635663917084864069175263416748948835535483336 \\ &4717309018125451550646500; \\ c(4,200) &=2.09455\dotsc\,. \end{aligned} \end{equation*} \notag $$
In Tables 26 we present some other results of computations in the same format as in [4]. All exact values computed are available on the webpage https://www.math.univ-toulouse.fr/~orevkov/tr.html.

Table 2

$n$Number of primitive triangulations of a $5\times n$ rectangle$c(5,n)$
12521.5954
21821321.7474
31828815201.8297
42089027667881.8802
52604205481449961.9155
63418164896255220321.9415
74644763856809356562401.9615
86458551594663713919476601.9773
99130369025134990418207027841.9902
1013065208497336167817891905138202.0008
1118875911658916512539040394323711722.0098
1227478484277212414619051763610781471682.0174
1340247583863108014277936023744662437146082.0240
1459247447360417186876229581918294710108471322.0298
1587579561995712611166902265987645011420884968602.0348
16129912159579165776352510956138594651762165301060802.0394
17193279021569720146452159319089306122189546163664646682.0434
18288288436487961179632386811809193620901579719205762139922.0470
$\vdots$$\vdots$$\vdots$
11518700706608364882730712710491937598381242505216572196 74626658766824095096227084981348969054292582022965697 97536209347455134357618461876316197344892595460029612 59669310339853198410108464789290118181041289819323068 31435995596306245022821112218622320544399050742600358 31426475886050757674088153732325783413307209633451618 73035677107305109076541667755690839416820326596 2.1044

Table 3

$n$Number of primitive triangulations of a $6\times n$ rectangle$c(6,n)$
19241.6419
228017081.7848
3122441844721.8617
4617562217429661.9088
53418164896255220321.9415
619992069347511330555181.9655
7121694099541419887071860521.9840
8760833363329475136555549189941.9987
94847725121672666884983996329181962.0107
1031315219598697701281384912878260659042.0206
11204437676119275998232172917694684494885482.0289
121345585503684000963645890647045368491317360242.0360
138915138987402468530383269504838128687912084420162.0421
1459387808248696685130595688923707759529337217433773542.0474
15397384566605094114342856423701539591155256038442585158602.0521
$\vdots$$\vdots$$\vdots$
50733088849377871573475229677373109896289395791929 288892292779893207423013116473882328714681504398 803902969400882970235141773360945092837017232937 18649959865340631279903635319082015514105847182.1023

Table 4

$n$Number of primitive triangulations of a $7\times n$ rectangle$c(7,n)$
134321.6778
2439368241.8134
38396606602681.8862
4187928962083870121.9307
54644763856809356562401.9615
6121694099541419887071860521.9840
73326338408441131037515979959202.0014
893693635175012088195304299672807082.0152
92696211097537325182524932578284131372722.0264
1078800099790205016140603947471701000930573002.0357
112330316428839061493866196473045629775863113725562.0435
1269536098303045180241255456746427705822741677605682602.0501
132089809948331032668557716536086803301598838540512759676122.0559
$\vdots$$\vdots$$\vdots$
2052066212145180734892042606757684021681422119 852336307301989140714761537366783840639832522.0813

Table 5

$n$Number of primitive triangulations of an $8\times n$ rectangle$c(8,b)$
1128701.7064
26986078161.8362
3585913812962561.9056
458315280224826297101.9480
56458551594663713919476601.9773
6760833363329475136555549189941.9987
793693635175012088195304299672807082.0152
811910648128826855397857137454009340443082.0282
91550233028202541336293688811781380767384621122.0388
10205273372387690323157963320071671029847454173440462.0476
1127538102329763517880812747863787333092362984269772038482.0550
123731191783577780617179480999800134602292060308057993985008542.0613
1350951326753537773696400958035190445392087069512323700346738258636 2.0668

Table 6

$n$Number of primitive triangulations of a $9\times n$ rectangle$c(9,n)$
1486201.7299
2112245984241.8547
341401067471782921.9214
418359333848129414533121.9621
59130369025134990418207027841.9902
64847725121672666884983996329181962.0107
72696211097537325182524932578284131372722.0264
81550233028202541336293688811781380767384621122.0388
9913765124094622356941511198970523445220062983109082.0489

2.3. Convexity conjecture for the numbers $f(m,n)$

The following conjecture is confirmed by all the computed exact values of the quantities $f(m,n)$ (we set $f(m,0)=1$ by convention).

Conjecture 2.5. The inequality

$$ \begin{equation*} f(m,n-1)f(m,n+1)\geqslant f(m,n)^2 \end{equation*} \notag $$
holds for all $m,n\geqslant 1$.

Proposition 2.6. If Conjecture 2.5 holds true, then $c_m\geqslant (n+1)c(m,n+1)-nc(m,n)$ for all $m,n\geqslant 1$. In particular, Conjecture 2.5 implies that

$$ \begin{equation*} c\geqslant c_{115}\geqslant 5c(115,5)-4c(115,4)=2.1684837\dotsc\,. \end{equation*} \notag $$

Proof. Set $d(m,n)=\log_2 f(m,n+1) - \log_2 f(m,n)$. Then Conjecture 2.5 implies that
$$ \begin{equation*} d(m,n)\leqslant d(m,n+1)\leqslant d(m,n+2)\leqslant\dotsb \, , \end{equation*} \notag $$
hence
$$ \begin{equation*} \log_2 f(m,n+k)-\log_2 f(m,n) \geqslant k d(m,n). \end{equation*} \notag $$
Dividing by $km$ and passing to the limit as $k\to\infty$ we obtain
$$ \begin{equation*} c_m\geqslant \frac{d(m,n)}{m}=(n+1)c(m,n+1)-nc(m,n). \end{equation*} \notag $$

The proposition is proved.

§ 3. The exact value of $c_2$ (proof of Theorem 1.1)

For $a,c\geqslant 0$, $a\equiv c\mod 2$, let $g^*_{a,c}$ denote the number of primitive lattice triangulations of the trapezoid $T(a,c)$ spanned by $(0,0)$, $(a,0)$, $(1,2)$ and $(1+c,2)$ (if ${a=0}$ or $c=0$, then $T(a,c)$ degenerates to a triangle). When $a\not\equiv c\mod 2$, we set $g^*_{a,c}=0$. We also set $g^*_{0,0}=1$. Let $G^*(x,z)$ be the generating function of the $g^*_{a,c}$:

$$ \begin{equation*} \begin{aligned} \, G^*(x,z) &= \sum_{a,c\geqslant0} g^*_{a,c}x^a z^c \\ &= 1+(x^2+xz+z^2)+(6x^4+10x^3 z+12 x^2 z^2+10 x z^3+6z^4)+\dotsb \end{aligned} \end{equation*} \notag $$
(here the definition of $g^*_{a,c}$ and $G^*$ is not the same as in Example 2.4). Let $g^*_n$ be the coefficient of $x^{2n}$ in the series $G^*(x,x)=\sum_{n\geqslant0} g^*_nx^{2n}$, that is,
$$ \begin{equation*} g^*_n=g^*_{0,2n}+g^*_{1,2n-1}+g^*_{2,2n-2}+\dots+g^*_{2n,0}. \end{equation*} \notag $$
Then Theorem 1.1 follows immediately from Lemmas 3.1 and 3.2 below.

Lemma 3.1. The equality

$$ \begin{equation*} \lim_{n\to\infty} f(2,n)^{1/n}=\lim_{n\to\infty} (g_n^*)^{1/n} \end{equation*} \notag $$
holds.

Proof. A $2\times(n-1)$ rectangle can be put in $T(n,n)$, hence $f(2,n-1) < g^*_n$. On the other hand the union of $T(a,c)$ with its image under the central symmetry with centre $(\frac12(a+c+1),1)$ is $T(a+c,a+c)$, and it can be put in a $2\times(a+c+1)$ rectangle. Hence $(g^*_{a,c})^2 < f(2,a+c+1)$. Therefore,
$$ \begin{equation*} \frac{g^*_n}{2n}=\sum_{a+c=2n}\frac{g^*_{a,c}}{2n} \leqslant\max_{a+c=2n}g^*_{a,c}\leqslant f(2,2n+1)^{1/2}\leqslant(g^*_{2n+2})^{1/2} \end{equation*} \notag $$
so that
$$ \begin{equation*} \frac1n\bigl(\log g^*_n-\log(2n)\bigr) \leqslant \frac1{2n}\log f(2,2n+1) \leqslant \frac1{2n}\log g^*_{2n+2}, \end{equation*} \notag $$
and the result follows because $\frac1n\log(2n)\to 0$.

The lemma is proved.

Lemma 3.2. The equality

$$ \begin{equation*} \lim_{n\to\infty}(g^*_n)^{1/n}=\alpha \end{equation*} \notag $$
holds, where $\alpha$ is as in Theorem 1.1.

Proof. For $a,c\geqslant 0$, $a\equiv c\mod 2$, let $g_{a,c}$ be the number of those primitive lattice triangulations of the trapezoid $T(a,c)$ that do not contain interior edges of the form $[(k,0),(l,2)]$, in other words, of the primitive lattice triangulations that agree with the subdivision of $T(a,c)$ into two triangles and two trapezoids shown in Figure 3 (left). If $a+c$ is odd, then we set $g_{a,c}=0$. By convention we set $g_{0,0}=0$. Let $G(x,z)=\sum_{a,c\geqslant 0} g_{a,c}x^a z^c$ be the corresponding generating function.

The edges of the form $[(k,0),(l,2)]$ of any primitive lattice triangulation cut $T(a,c)$ into smaller trapezoids. These can be transformed into trapezoids $T(a_i,c_i)$ such that $\sum a_i = a$ and $\sum c_i = c$ by uniquely determined lattice automorphisms of the form $(x,y)\mapsto(x+p_iy+q_i,y)$, where $p_i,q_i\in\mathbb{Z}$ (see Figure 3). Hence

$$ \begin{equation*} g^*_{a,c}= \sum_{\substack{a_1+\dots+a_k=a \\ c_1+\dots+c_k=c}}\, \prod_{j=1}^k g_{a_j,c_j}, \end{equation*} \notag $$
and thus,
$$ \begin{equation} G^*(x,z)=\frac1{1-G(x,z)}. \end{equation} \tag{2} $$

It is easy to see (cf. (1)) that the numbers of primitive lattice triangulations of the narrow (that is, of width 1) trapezoids in Figure 3 are binomial coefficients, hence

$$ \begin{equation*} G(x,z)=(x^2+xz+z^2)+(5x^4+8x^3z+9x^2z^2+8xz^3+5z^4)+\dotsb. \end{equation*} \notag $$

One can also check that $g_{a,c}=f_{a,(a+c)/2-1,c}$ where the $f_{a,b,c}=f(S_{a,b,c})$ are the numbers discussed in Example 2.3. Hence (cf. Example 2.4)

$$ \begin{equation*} \begin{aligned} \, G(x,z) &=\sum_{a,c} f_{a,(a+c)/2-1,c}x^az^c =\operatorname{coef}_{u^0} \biggl[\sum_{a,b,c}f_{a,b,c}x^a u^{2b-a-c+2}z^c\biggr] \\ &=\operatorname{coef}_{u^0}\biggl[u^2\sum_{a,b,c}f_{a,b,c}(xu^{-1})^a u^{2b}(zu^{-1})^c\biggr] =\operatorname{coef}_{u^{-1}}[u F(xu^{-1},u^2,zu^{-1})]. \end{aligned} \end{equation*} \notag $$
Since the function $1/(1-x-y-z+xz)=1/\bigl((1-x)(1-z)-y\bigr)$ is analytic in the domain $\max\bigl(|x|,|2y|,|z|\bigr)<1/2$, its power series $\sum f_{a,b,c}x^ay^bz^c$ (see Example 2.3) converges to it in this domain. Therefore, for $0<\varepsilon\ll r<1/2$ the Laurent series of $F(x/u,u^2,z/u)$ converges in the domain $\max(|x|,|z|)<\varepsilon$, $r-\varepsilon<|u|<r+\varepsilon$. Hence for $x$ small enough we have
$$ \begin{equation*} G(x,x)=\operatorname{coef}_{u^{-1}}\biggl[F\biggl(\frac xu,u^2,\frac xu\biggr)\biggr] =\frac{1}{2\pi i}\oint_{|u|=r}\frac{u\,du}{(1-x/u)^2-u^2} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, \frac{u}{(1-x/u)^2-u^2} &= -\frac{u}{2(u^2+u-x)}-\frac{u}{2(u^2-u+x)} \\ &=\sum_{j=1}^2\frac 1{2(u_j^+-u_j^-)}\biggl(\frac{u_j^+}{u-u_j^+}+\frac{u_j^-}{u-u_j^-}\biggr), \end{aligned} \end{equation*} \notag $$
where, for $|x|$ small enough,
$$ \begin{equation*} u_1^\pm=-\frac12(1\pm\sqrt{1+4x})\quad\text{and} \quad u_2^\pm=\frac12(1\pm\sqrt{1-4x}), \qquad |u_j^+|>r, \quad |u_j^-|<r. \end{equation*} \notag $$
Thus,
$$ \begin{equation*} G(x,x)=\sum_{j=1}^2\operatorname*{Res}_{u=u_j^-}(\dots) =\sum_{j=1}^2\frac{u_j^-}{2(u_j^+-u_j^-)} =\frac1{4\sqrt{1-4x}}+\frac1{4\sqrt{1+4x}}-\frac12. \end{equation*} \notag $$
The graph of the function $y = G(x,x)$ sits in the algebraic curve
$$ \begin{equation*} (2y+1)^2(16x^2-1)\bigl(4x^2+(y^2+y)(16x^2-1)\bigr)+x^2=0. \end{equation*} \notag $$
By (2) the poles of $G^*(x,x)$ are the $x$-coordinates of the points of intersection of this curve with the line $y=1$, that is, the roots of the polynomial $5184x^4 - 611 x^2 + 18$ (the smallest of which are $\pm\sqrt{1/\alpha}$), and the branch points are $\pm1/4$. Hence the radius of convergence of the series $G^*(x,x)=\sum g^*_n x^{2n}$ is $\sqrt{1/\alpha}$, so that
$$ \begin{equation*} \lim_{n\to\infty} (g^*_n)^{1/n}=\alpha. \end{equation*} \notag $$

Lemma 3.2 is proved.

§ 4. Computation of $c_3$ (proof of Theorem 1.2)

4.1. Preparation

For $a,d\geqslant0$ such that $a\not\equiv d+1\mod3$, let $h^*_{a,d}$ be the number of primitive lattice triangulations of the trapezoid $T_3(a,d)$ spanned by $(0,0)$, $(1,3)$, $(1+d,3)$ and $(a,3)$. We set $h^*_{0,0}=1$ and $h^*_{a,d}=0$ for $a\equiv d+1\mod 3$, and we consider the generating function

$$ \begin{equation*} H^*(x)=\sum_n h^*_n x^n=\sum_{a,d\geqslant 0}h^*_{a,d}x^{a+d} =1+x+3x^2+19x^3+125 x^4 +\dotsb. \end{equation*} \notag $$
Similarly to the beginning of the proof of Lemma 3.2, we define $h_{a,d}$ as the number of the triangulations of $T_3(a,d)$ that do not have edges of the form $[(k,0),(l,3)]$, and we consider the generating function
$$ \begin{equation*} H(x)=\sum_n h_n x^n=\sum_{a,d\geqslant 0}h_{a,d} x^{a+d}=x+2x^2+14 x^3+86 x^4+712 x^5 +\dotsb. \end{equation*} \notag $$
These functions satisfy a relation similar to (2) specialized for $x=z$:
$$ \begin{equation*} H^*(x)=\frac{1}{1-H(x)}. \end{equation*} \notag $$
Indeed, the edges of the form $[(k,0),(l,3)]$ cut $T_3(a,d)$ into smaller trapezoids. Each of these can be mapped to a standard trapezoid by a unique lattice automorphism of the form $(x,y)\mapsto(x+py+q,y)$ or $(x,y)\mapsto(x+py+q,3-y)$, where $p,q\in\mathbb{Z}$ (in contrast to § 2, here the upper and lower horizontal sides of the trapezoids can be interchanged, so we do not have (2) for the two-variable generating functions).

We illustrate the relation

$$ \begin{equation*} h^*_3=h^*_{03}+h^*_{12}+h^*_{30} =h_{01}^3+2h_{01}(h_{11}+h_{20})+(h_{03}+h_{12}+h_{30}) =h_1^3+2 h_1 h_2+h_3 \end{equation*} \notag $$
in Figure 4.

Similarly to Lemma 3.1 we have

$$ \begin{equation*} \lim_n f(3,n)^{1/n}=\lim_n(h^*_{2n})^{1/n}=\frac{1}{\beta^{2}}, \end{equation*} \notag $$
where $\beta$ is the first real root of the equation $H(x)=1$; hence
$$ \begin{equation*} c_3=-\frac23\log_2\beta. \end{equation*} \notag $$

4.2. Recurrence relations

Using the notation introduced in § 2, set $m=3$, $\varphi_0(x)=\frac13x-1$ and

$$ \begin{equation*} \begin{gathered} \, F(x,y,z,w)=\sum_{a,b,c,d} f_{a,b,c,d}x^ay^bz^cw^d, \\ G_1(x,z,w)=\sum_{a,c,d} g^{(1)}_{a,c,d}x^a z^c w^d, \qquad G_2(x,y,w)=\sum_{a,b,d} g^{(2)}_{a,b,d}x^a y^b w^d, \\ H_k(x,w)=\sum_{a,d}h^{(k)}_{a,d}x^a w^d, \qquad k=1,2, \end{gathered} \end{equation*} \notag $$
where all coefficients are of the form $f(S)$ (see § 2) for the $\varphi_0$-admissible shapes in Figure 5, where $(0,a)$, $(1,b)$, $(2,c)$ and $(3,d)$ (if present) are the coordinates of integral points on the upper part of the boundary of $S$. The lower corners of $S$ are at the points $(0,-1)$ and $(3,0)$. If the congruences in Figure 5 are not satisfied, then the corresponding numbers are zero. If $\min(a+1,b,c,d)<0$, then they are also zero (this case does not correspond to any $\varphi_0$-admissible shape). We also set $h_{-1,0}^{(2)}=0$ by convention (the case when $S$ degenerates to a line segment).

In terms of the generating functions the recurrence relations in Lemma 2.2 read (cf. Examples 2.3 and 2.4)

$$ \begin{equation*} \begin{gathered} \, \begin{split} & F(x,y,z,w)(1-x-y-z-w+xz+yw+xw) \\ &\qquad =y^{1/2}(1-w)G_1(xy^{1/2},y^{1/2}z,w)+z^{1/2}(1-x)G_2(x,yz^{1/2},z^{1/2}w), \end{split} \\ \begin{split} G_1(x,z,w)(1-w) &=\operatorname{coef}_{u^{-1}}\biggl[F\biggl(\frac xu,u^2,\frac zu,w\biggr)(1-w)\biggr]+x^{-1}, \\ G_2(x,y,w)(1-x) &=\operatorname{coef}_{u^{-1}}\biggl[F\biggl(x,\frac yu,u^2,\frac wu\biggr)(1-x)\biggr] \end{split} \end{gathered} \end{equation*} \notag $$
(the asymmetry between $G_1$ and $G_2$ is caused by the asymmetry of $\varphi_0$),
$$ \begin{equation*} H_1(x,w) =\operatorname{coef}_{u^{-1}}\biggl[ G_1\biggl(\frac xu,u^3,\frac{w}{u^2}\biggr)\biggr] \end{equation*} \notag $$
and
$$ \begin{equation*} H_2(x,w) =\operatorname{coef}_{u^{-1}}\biggl[ G_2\biggl(\frac{x}{u^2},u^3,\frac wu\biggr)\biggr]. \end{equation*} \notag $$
Notice that in this subsection, by generating functions we mean formal series. Consider the symmetrized generating functions
$$ \begin{equation*} \begin{gathered} \, \widetilde F(x,y,z,w) = F(x,y,z,w)+F(w,z,y,x), \\ \widetilde G(x,z,w) = G_1(x,z,w)+G_2(w,z,x) \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \widetilde H(x,w) = H_1(x,w)+H_2(w,x). \end{equation*} \notag $$
The above relations for $F$, $G_1$, $G_2$, $H_1$ and $H_2$ imply immediately that
$$ \begin{equation} \begin{split} &\widetilde F(x,y,z,w)(1-x-y-z-w+xz+yw+xw ) \\ &\qquad =y^{1/2}(1-w)\widetilde G(xy^{1/2},y^{1/2}z,w) +z^{1/2}(1-x)\widetilde G(x,yz^{1/2},z^{1/2}w), \end{split} \end{equation} \tag{3} $$
$$ \begin{equation} \widetilde G(x,z,w)(1-w)=\operatorname{coef}_{u^{-1}} \biggl[\widetilde F\biggl(\frac xu,u^2,\frac zu,w\biggr)(1-w)\biggr]+x^{-1} \end{equation} \tag{4} $$
and
$$ \begin{equation} \widetilde H(x,w)=\operatorname{coef}_{u^{-1}} \biggl[\widetilde G\biggl(\frac xu,u^3,\frac{w}{u^2}\biggr)\biggr]. \end{equation} \tag{5} $$

4.3. The equation

We are going to obtain an equation for $\widetilde G(xt^{-1/2},t^{3/2},t^{-1}x)$ by expressing $\widetilde F$ in terms of $\widetilde G$ from (3) and plugging it into (4). To do this we must divide power series by polynomials. However, when some variables appear in powers ranging from $-\infty$ to $+\infty$, the meaning of this division should be clarified. To illustrate the possible ambiguity consider the expression $\operatorname{coef}_{u^{-1}}\bigl[1/(x-uy)\bigr]$. It can be understood either as

$$ \begin{equation*} \operatorname{coef}_{u^{-1}}\biggl[ \frac{x^{-1}}{1-uyx^{-1}}\biggr] =\frac1x\operatorname{coef}_{u^{-1}}\biggl[1+\frac{uy}x+\frac{u^2y^2}{x^2}+\dotsb\biggr] =0, \end{equation*} \notag $$
or as
$$ \begin{equation*} \operatorname{coef}_{u^{-1}}\biggl[ -\frac{(uy)^{-1}}{1-x(uy)^{-1}}\biggr] =-\operatorname{coef}_{u^{-1}}\biggl[\frac1{uy}\biggl(1+\frac{x}{uy}+\frac{x^2}{u^2y^2} +\dotsb\biggr)\biggr] =-\frac1y. \end{equation*} \notag $$
To avoid this kind of ambiguity, we introduce a new formal variable $q$ and consider the formal series
$$ \begin{equation*} \begin{gathered} \, F_q(x,y,z,w)=F(xq,yq^2,zq^2,wq), \\ G_{1,q}(x,z,w)=G_1(xq^2,zq^3,wq), \\ G_{2,q}(x,y,w)=G_2(xq,yq^3,wq^2) \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} H_{k,q}(x,w)=H_k(xq^3,wq^3), \qquad k=1,2; \end{equation*} \notag $$
all generating functions will be treated as elements of the ring
$$ \begin{equation*} \mathbb Z\bigl[x^{\pm1},y^{\pm1/2},z^{\pm1/2},w^{\pm1},u^{\pm1/2},t^{\pm1/2}\bigr]((q)) \end{equation*} \notag $$
of formal power series in $q$ (maybe starting with a negative power) whose coefficients are Laurent polynomials in $x,y^{1/2},\dots$ .

The geometric meaning of an exponent of $q$ is twice the signed area of the $\varphi_0$-admissible shape corresponding to the monomial under consideration, that is, $\displaystyle2\int_0^3\varphi(x)\,dx$ where the graph of $\varphi$ is the upper boundary of the shape. One can easily check by hand that

$$ \begin{equation*} \begin{aligned} \, F_q &=(xq)^{-1}+(1+x^{-1}w)+(x+w+x^{-1}y+x^{-1}z+x^{-1}w^2)q+\dotsb, \\ G_{1,q} &=x^{-1}q^{-2}+w(xq)^{-1}+w^2x^{-1}+w^3x^{-1}q+(x+w^4x^{-1})q^2+\dotsb, \\ G_{2,q} &=x^{-1}wq+(w+yx^{-1})q^2+(wx+2y)q^3+(wx^2+4xy)q^4+\dotsb, \\ H_{1,q} &=wx^{-1}+xq^3+4w^2 q^6+(30wx^2+24 w^4 x^{-1})q^9+\dotsb, \\ H_{2,q} &=wq^3+5(x^2+w^3x^{-1})q^6+ 32w^2xq^9+\dotsb. \end{aligned} \end{equation*} \notag $$

Further, we define $\widetilde F_q$, $\widetilde G_q$ and $\widetilde H_q$ by the same formulae as $\widetilde F $, $\widetilde G $ and $\widetilde H $ in § 4.2, adding the subscript $q$ everywhere. For example,

$$ \begin{equation*} \widetilde G_q(x,z,w)=\frac{1}{xq^2}+\frac w{xq}+\frac{w^2}x +\biggl(\frac{w^3}x+\frac xw\biggr)q +\biggl(2x+\frac{w^4}x+\frac zw\biggr)q^2+\dotsb. \end{equation*} \notag $$
Then relations (3)(5) take the form
$$ \begin{equation} \widetilde F_q=\frac{qy^{1/2}(1-wq)\widetilde G_q(xy^{1/2},y^{1/2}z,w) +qz^{1/2}(1-xq)\widetilde G_q(wz^{1/2},z^{1/2}y,x)} {1-xq-yq^2-zq^2-wq+xzq^3+ywq^3+xwq^2 }, \end{equation} \tag{6} $$
$$ \begin{equation} \widetilde G_q(x,z,w) =\operatorname{coef}_{u^{-1}}\biggl[q\widetilde F_q\biggl(\frac xu,u^2,\frac zu,w\biggr)\biggr] +\frac1{x(1-wq)q^2}, \end{equation} \tag{7} $$
$$ \begin{equation} \widetilde H_q(x,w)=\operatorname{coef}_{u^{-1}} \biggl[q\widetilde G_q\biggl(\frac xu,u^3,\frac{w}{u^2}\biggr)\biggr]. \end{equation} \tag{8} $$

Let us set

$$ \begin{equation*} \begin{aligned} \, g_q(x,t) &= t^{1/2}x^2q^2 \widetilde G_q(x^2t^{-1/2},x^3t^{3/2},xt^{-1}) \\ &=t+xq+t^{-1}x^2q^2+(t^{-2}+t)x^3q^3+(t^{-3}+2+t^3)x^4 q^4+\dotsb. \end{aligned} \end{equation*} \notag $$
The parity condition on the indices of nonzero coefficients of $G_1$ and $G_2$ (see Figure 5) ensures that the series $g_q(x,t)$ does not contain fractional powers. Moreover, $x$ and $q$ have the same power in each monomial of $g_q$, thus $g_q(x,t)=g(xq,t)$ for some $g(x,t)\in\mathbb{Z}[t^{\pm1}]((x))$.

Plugging (3) into (4), denoting the denominator in (6) by $Q_q(x,y,z,w)$ and observing that

$$ \begin{equation} \operatorname{coef}_{u^{-1}}\bigl[\mathcal F(x,t,u)\bigr] =\operatorname{coef}_{u^{-1}}\bigl[xt^{-1/2}\mathcal F(x,t,uxt^{-1/2})\bigr] \end{equation} \tag{9} $$
for any formal Laurent series in $u$ we obtain
$$ \begin{equation*} \begin{aligned} \, g_q(x,t) &\stackrel{(7)}{=} \operatorname{coef}_{u^{-1}}\biggl[t^{1/2} x^2 q^3 \widetilde F_q\biggl(\frac{x^2}{ut^{1/2}},u^2,\frac{x^3t^{3/2}}{u},\frac{x}{t}\biggr)\biggr] +\frac{t^2}{t-xq} \\ &\stackrel{(9)}{=} \operatorname{coef}_{u^{-1}}\biggl[x^3 q^3 \widetilde F_q\biggl(\frac{x}{u},\frac{x^2u^2}{t},\frac{x^2t^2}{u},\frac{x}{t}\biggr)\biggr] +\frac{t^2}{t-xq} \\ &\stackrel{(6)}{=} x^2q^2\operatorname{coef}_{u^{-1}}\biggl[\frac{\frac{u}{t}\bigl(1-\frac{xq}{t}\bigr)g_q(x,t) +\frac{t}{u}\bigl(1-\frac{xq}{u}\bigr)g_q(x,u)} {Q_q\bigl(x/u,\,x^2u^2/t,\,x^2t^2/u,\,x/t\bigr)}\biggr] +\frac{t^2}{t-xq} \\ &= x^2q^2\operatorname{coef}_{u^{-1}}\biggl[\frac{u^3(t-xq)g_q(x,t)+t^3(u-xq)g_q(x,u)}{P(xq,t,u)} \biggr]+\frac{t^2}{t-xq}, \end{aligned} \end{equation*} \notag $$
where
$$ \begin{equation} P(x,t,u)=u^2t^2-(u+t)utx +(1-t^3-u^3)utx^2+ (t^4+u^4)x^3. \end{equation} \tag{10} $$

We see that the variables $x$ and $q$ are ‘synchronized’ on the right-hand side of the equation obtained: they occur to the same power in each monomial of each power series in this expression. Hence we obtain the following identity in the ring $\mathbb{Z}[t^{\pm1},u^{\pm1}]((x))$:

$$ \begin{equation} g(x,t)\Psi(x,t)=\frac{t^2}{t-x}+\operatorname{coef}_{u^{-1}} \biggl[\frac{t^3x^2(u-x)g(x,u)}{P(x,t,u)}\biggr], \end{equation} \tag{11} $$
where
$$ \begin{equation*} \Psi(x,t)=1-x^2(t-x)\Phi(x,t) \quad\text{and}\quad \Phi(x,t)=\operatorname{coef}_{u^{-1}}\biggl[\frac{u^3}{P(x,t,u)}\biggr]. \end{equation*} \notag $$
Here are several initial terms of these series:1
$$ \begin{equation} \begin{aligned} \, \notag \Phi(x,t) &=t^{-2}x^2+(t^{-3}+1)x^3+(t^{-4}+2t^{-1}+t^2)x^4 +(t^{-5}+3t^{-2}+3t)x^6+\dotsb, \\ \Psi(x,t) &=1-t^{-1}x^4-tx^5-(1+t^3)x^6-(t^{-1}+2t^2)x^7-(6t^{-2}+3t)x^8-\dotsb. \end{aligned} \end{equation} \tag{12} $$

Having found $g$ from (11), we can compute $\widetilde H(x,x)$. Indeed, by (5) we have

$$ \begin{equation*} x\widetilde H_q(x^3,x^3) =\operatorname{coef}_{t^0}\biggl[tx\widetilde G_q \biggl(\frac{x^3}t,t^3,\frac{x^3}{t^2}\biggr)\biggr] =\operatorname{coef}_{t^0}\biggl[t^{1/2}x\widetilde G_q \biggl(\frac{x^3}{t^{1/2}},t^{3/2},\frac{x^3}{t}\biggr)\biggr]. \end{equation*} \notag $$
Replacing $t$ by $x^2t$ (cf. (9)) and setting $q=1$ we obtain
$$ \begin{equation} x \widetilde H(x^3,x^3)=\operatorname{coef}_{t^{0}}\bigl[g(x,t)\bigr]. \end{equation} \tag{13} $$

4.4. Computation

In this subsection we study the analytic functions defined by the series discussed in the previous subsection.

By § 4.1 we need to find the smallest positive pole of $H^*(x)$, that is the smallest positive zero $\beta$ of $1-H(x)$. One can check that

$$ \begin{equation} H(x)=x\widetilde H(x,x). \end{equation} \tag{14} $$
Being the sum of a power series with positive coefficients, the function $x\widetilde H(x,x)$ is increasing for $x>0$, thus it is enough to know how to compute the value of $\widetilde H(x,x)$ with any prescribed precision for any fixed $x$ in an interval containing $\beta$. By (13) this can be done by integrating the function $g(x^{1/3},t)$ numerically over a suitable contour $\Gamma_x$ (cf. the proof of Lemma 3.2). Thus we need to be able to compute $g(x,t)$ for any $x\in[0,x_0^+]$ and $t\in\Gamma_x$, for some $x_0^+ > x_0=\beta^{1/3}$. This can be done because for fixed $x$, after replacing $\operatorname{coef}_{u^{-1}}[\dots]$ by $\displaystyle \frac1{2\pi i}\int_{\Gamma_x}(\dots)\,du$, identity (11) becomes a Fredholm equation for the restriction of the function $g$ to $\Gamma_x$. Now we go over to more detailed explanations.

Let

$$ \begin{equation*} \Gamma =\biggl\{(x,t,u)\in\mathbb R\times\mathbb C^2\Bigm| 0<x<\frac 12,\, |t|=|u|=1\biggr\} \end{equation*} \notag $$
and
$$ \begin{equation*} \Gamma' =\biggl\{(x,t)\in\mathbb R\times\mathbb C\Bigm| 0<x<\frac 12,\, |t|=1\biggr\}. \end{equation*} \notag $$

Lemma 4.1. The polynomial $P(x,t,u)$ defined in (10) does not vanish on $\Gamma$. For any fixed point $(x,t)\in\Gamma'$ the polynomial $P(x,t,u)$ has two simple roots $u_k(x,t)$, $k=1,2$, in the unit disc $|u|<1$ and two simple roots outside it.

Proof. The first statement can be checked using any software for symbolic computations. This can be done, for example, as follows. Let $S^1$ be the unit circle in $\mathbb{C}$. Then $\Gamma=(0,1/2)\times S^1\times S^1$. We can identify $S^1$ with $\mathbb{RP}^1$ by means of some rational parametrization. Then $\operatorname{Re} P$ and $\operatorname{Im} P$ become real rational functions on the variety $\Gamma$ and, by computing resultants, the discriminant and so on, one can check that the real algebraic curve given by the equations $\operatorname{Re} P=\operatorname{Im} P=0$ does not intersect the layer $0<r<1/2$. More precisely, let $p(x,T,U)$ and $q(x,T,U)$ be real polynomials such that
$$ \begin{equation*} P(x,\zeta(T),\zeta(U))=\frac{p(x,T,U)+iq(x,T,U)}{(i+T)^4(i+U)^4} \quad\text{and}\quad \zeta(X)=\frac{i-X}{i+X}. \end{equation*} \notag $$
Note that $\zeta(\mathbb{R})=S^1\setminus\{-1\}$, hence $(x,T,U)$ are coordinates on the affine chart $\Gamma\setminus\{(t+1)(u+1)=0\}$ of $\Gamma$. The projection of the real algebraic curve $\Gamma\cap\{P=0\}$ onto the plane $(x,T)$ is given by the equation $R(x,T)=0$, where $R(x,T)$ is the resultant of $p$ and $q$ with respect to $U$. To prove that the curve $R(x,T)=0$ does not have real points such that $0<x<1/2$ we compute the real roots of $D(x)=0$ on this interval, where $D(x)$ is the discriminant of $R$ with respect to $T$. Then we check that the equations $R(x_k,T)=0$ do not have real roots for $0<x_1<\dots<x_{2n+1}<1/2$, where the points $x_k$ with even $k$ are the real roots of $D(x)$ on the interval $0<x<1/2$. This computation shows that $P(x,u,t)\ne0$ for $(x,u,t)\in\Gamma$ and $(t+1)(u+1)\ne0$. Then we check that $P(x,\zeta(T),-1)\ne0$, $P(x,-1,\zeta(U))\ne0$ and $P(x,-1,-1)\ne0$ for $0<x<1/2$ and $T\in\mathbb{R}$.

One can check similarly that for any fixed $(x,t)\in\Gamma'$ the discriminant of $P$ with respect to the variable $u$ does not vanish, hence for any fixed $(x,t)\in\Gamma'$ all the four roots of $P$ (viewed as a polynomial in $u$) are pairwise distinct.

Therefore, the number of roots of $P$ in the unit disc $|u|<1$ is constant. Thus, to prove the second statement it is enough to check it for some values of $x$ and $t$, for example, for $t=1$ and some very small $x$.

The lemma is proved.

Lemma 4.2. (a) The formal power series $1/P(x,t,u)\in\mathbb{Z}[t^{\pm1},u^{\pm1}]((x))$ converges to the function $1/P(x,t,u)$ in a neighbourhood of $\Gamma\cap\{x< 1/4\}$.

(b) The formal power series $\Phi(x,t)\in\mathbb{Z}[t^{\pm1}]((x))$ converges to an analytic function (which we also denote by $\Phi(x,t)$) in a neighbourhood of $\Gamma'\cap\{x<1/4\}$. The function $\Phi(x,t)$ admits an analytic extension to a neighbourhood of $\Gamma'$, which is defined by the Cauchy integral

$$ \begin{equation} \Phi(x,t)=\frac{1}{2\pi i}\oint_{|u|=1}\frac{u^3\,du}{P(x,t,u)} =\sum_{k=1}^2\frac{u_k(x,t)^3}{P'_u(x,t,u_k(x,t))}, \end{equation} \tag{15} $$
where $u_1(x,t)$ and $u_2(x,t)$ are the roots of $P$ in the unit disc $|u|<1$; see Lemma 4.1.

Proof. The power series $1/P(x,t,u)$ involved in the definition of $\Phi(x,t)$ is a power series expansion with respect to $x$, hence $1/P=a_0^{-1}(1+X+X^2+\cdots)$ where $X=(a_1+a_2+a_3)/a_0$ and $a_k=x^k\operatorname{coef}_{x^k}[P]$. If $(x,t,u)\in\Gamma$, then $|a_0|=1$, $|a_1|\leqslant 2x$, $|a_2|\leqslant 3x^2$ and $|a_3|\leqslant 2x^3$, and thus $|X|\leqslant 2x+3x^2+2x^3$. Therefore, $|X|<1$ for $x<1/4$, which shows the convergence of $1/P$ in the required domain. This fact in combination with Lemma 4.1 implies all the other assertions of the lemma.

The lemma is proved.

The function Psi in the Mathematica system that is presented in Figure 7 computes $\Psi(x,t)$ for $(x,t)\in\Gamma'$ with any prescribed precision.

Notice that one of the functions $u_1(x,t)$ or $u_2(x,t)$ has a branch point at $(x,t)=(1/2,1)$, and hence the functions $\Phi$ and $\Psi$ are also branched at this point. The Laurent-Puiseux expansion of $\Psi(x,1)$ in powers of $s=\sqrt{1/2-x}$ is

$$ \begin{equation*} \begin{aligned} \, \Psi(x,1) &=-\frac{1}{4\sqrt6}\,s^{-1}+\frac{12-\sqrt2}{8}-\frac{3}{8\sqrt6}\,s -\frac{3}{8\sqrt2}\,s^2 \\ &\qquad+\frac{103}{96\sqrt6}\,s^3-\frac{87}{32\sqrt2}\,s^4 +\frac{2635}{192\sqrt6}\,s^5+\dotsb. \end{aligned} \end{equation*} \notag $$

Let $x_0^-={16}/{33}$ and $x_0^+={17}/{35}$. We will see below that $x_0\in[x_0^-,x_0^+]$; in fact, $x_0^{\pm}$ are given by initial segments of the continued fraction of $x_0$.

Using the expansion of $\Psi$ at $(1/2,1)$ and computing the values of $\Psi(x,t)$ on a sufficiently dense grid on $\Gamma'$ (using the program presented in Figure 7) one can check that $\Psi$ does not vanish on $\Gamma'\cap\{x<x_0^+\}$ and

$$ \begin{equation} \min_{0\leqslant x\leqslant x_0^+,\,|t|=1}|\Psi(x,t)| =\min_{0\leqslant x\leqslant x_0^+,\,|t|=1}\operatorname{Re}\Psi(x,t) =\Psi(x_0^+,1)=0.44768\dots; \end{equation} \tag{16} $$
see the level curves of $\operatorname{Re}\Psi$ in Figure 6; we omit the details of the error estimate.

Applying Lemma 5.2 for an appropriately chosen $h$ to the function $|P(x/4,e^{i\tau},e^{i\theta})|^2$ we find that

$$ \begin{equation} \min_{0\leqslant x\leqslant x_0^+,\,|t|=|u|=1}|P|=P(x_0^+,1,1)=0.02183\dots \end{equation} \tag{17} $$
(here we have rescaled $x$ to equilibrate the partial derivatives). The computation can be accelerate by choosing different grids in different zones of $\Gamma$. In our computations the grid step varied from $h=1/300$ near the point of minimum to ${h=1/20}$ far away from it. To estimate the error we used evident coarse bounds for the fourth derivatives and, using them, computed finer upper bounds for the second derivatives in each zone, using Lemma 5.2 again.

Lemma 4.3. The formal series $g(x,t)$ (introduced in § 4.2) converges in some neighbourhood of $\Gamma'\cap\{|x|<2^{-3/2}\}$.

Proof. By Anclin’s theorem [1] the number of primitive lattice triangulations of a lattice polygon $\Pi$ is bounded above by $2^N$ for $N=\#\bigl(\Pi\cap(\mathbb Z^2\setminus\frac12\mathbb Z^2)\bigr)$, and it is easy to derive from Pick’s formula that $N<3\operatorname{Area}(\Pi)-3/2$. The area of the shape corresponding to $g^{(k)}_{a,c,d}$ is $(2a+3c+d+3)/2$. Hence $\widetilde g_{a,c,d} < c_0 2^{3(2a+3c+d)/2}$ for some constant $c_0$, and for $|t|=1$ we obtain
$$ \begin{equation*} \begin{aligned} \, |g(x,t)| &\leqslant x^2\sum_{a,c,d}|\widetilde g_{a,c,d}x^{2a} x^{3c} x^d| \\ &\leqslant c_0x^2\sum_{a,c,d}|2^{3(2a+3c+d)/2}x^{2a+3c+d}| =c_0x^2\sum_n 2^{3/2n}A_n x^n, \end{aligned} \end{equation*} \notag $$
where $A_n=\#\bigl\{(a,c,d)\in\mathbb Z_+^3\mid 2a+3c+d=n\bigr\}$. Since $A_n$ is bounded by a polynomial function of $n$, the series converges for $x<2^{-3/2}$.

The lemma is proved.

Lemmas 4.2 and 4.3 in combination with (11) and (16) imply that the function $g(x,t)$ is analytic in a neighbourhood of $\Gamma'\cap\{x<2^{-3/2}\}$ and satisfies there the condition

$$ \begin{equation} g(x,t)=\frac{t^2}{(t-x)\Psi(x,t)} +\frac{1}{2\pi i}\oint_{|u|=1}\frac{x^2t^3(u-x)g(x,u)\,du}{P(x,t,u)\Psi(x,t)}. \end{equation} \tag{18} $$
For any fixed $x$ this is Fredholm’s equation of the second kind for $g(x,t)$ as a function of $t$.

Lemma 4.4. The function $g(x,t)$ extends analytically to a neighbourhood of ${\Gamma'\cap\{x<x_0^+\}}$ and satisfies equation (18) in this domain.

Proof. Let us rewrite (18) in a more conventional form:
$$ \begin{equation} \varphi_g(x,\tau)=f(x,\tau) +\int_0^1 K(x,\tau,\theta)\varphi_g(x,\theta)\,d\theta, \end{equation} \tag{19} $$
where we set $t=e^{2\pi i\tau}$, $u=e^{2\pi i\theta}$ and
$$ \begin{equation*} \varphi_g(x,\tau)=g(x,t), \qquad f(x,\tau)=\frac{t^2}{(t-x)\Psi(x,t)}, \qquad K(x,\tau,\theta)=\frac{x^2t^3u(u-x)}{P(x,t,u)\Psi(x,t)}. \end{equation*} \notag $$

As we have already pointed out, $g$ satisfies (18) and thus $\varphi_g$ satisfies (19) for small $x$. Thus, thanks to the identity theorem for analytic functions, it is enough to show that a unique solution of (19) exists for any $x\in[0,x_0^+]$ and it is analytic with respect to $(x,\tau)$. Hence, by Lemma 5.6 it suffices to show that $1$ is not an eigenvalue of $\mathcal K_x$ for any $x\in[0,x_0^+]$, where $\mathcal K_x\colon\mathcal C[0,1]\to\mathcal C[0,1]$ is the Fredholm integral operator that takes $\varphi(\tau)$ to $\displaystyle\psi(\tau)=\int_0^1 K(x,\tau,\theta)\varphi(\theta)\,d\theta$. The latter fact, in its turn, follows from the bound

$$ \begin{equation*} \max_{0\leqslant x\leqslant x_0^+}\mathcal N_2(x)=\mathcal N_2(x_0^+) =0.88525, \end{equation*} \notag $$
where $\displaystyle\mathcal N_2(x)=\int_{[0,1]^2}|K(x,\tau,\theta)|^2\,d\tau\,d\theta$. This bound is computed using numerical integration. To estimate the approximation error one needs upper bounds for partial derivatives of $K$. They can easily be obtained using the lower bounds (16) and (17) for $|\Psi|$ and $|P|$ and the upper bounds for the derivatives of $\Psi$ that are obtained from its integral representation in (15). As upper bounds for the derivatives of polynomials involved in the definition of $K$ we can take just the sums of upper bounds for monomials.

The lemma is proved.

Replacing the integrals by integral sums, equation (18) can be solved with any prescribed precision. Then, in view of (13) and (14) we can numerically compute $H(x)$ using the Cauchy integral

$$ \begin{equation} H(x^3)=\frac{x^2}{2\pi i}\oint_{|t|=1}\frac{g(x,t)\,dt}{t} =x^2\int_0^1 \varphi_g(x,\tau)\,d\tau \end{equation} \tag{20} $$
(recall that $\varphi_g(x,\tau):=g(x,e^{2\pi i\tau})$; see (19)). We can summarize the content of this section as follows (recall that $f(m,n)$ is the number of primitive lattice triangulations of an $m\times n$ rectangle).

Proposition 4.5. The relation

$$ \begin{equation*} \lim_{n\to\infty}f(3,n)^{1/n}=\frac{1}{x_0^2} \end{equation*} \notag $$
holds, where:

In Figure 7 we present a Mathematica function H that computes $H(x)$ with any prescribed precision. The approximation error can be estimated using Lemma 5.4. One can check that the functions $P(x,t,u)$ and $\Psi(x,t)$ do not vanish for $x<x_0^+$, $|u|=1$ and ${10}/{13}<|t|<{13}/{10}$.

In Figures 8 and 9 we show the image of the annulus ${10}/{13}<|t|<{13}/{10}$ under the map $t\mapsto\Psi(x_0,t)$. Thus we can apply the error estimate (27) for $r=10/13$ and therefore for $a=-\log r/(2\pi)=0.04176$.

When estimating the error of $H(x)$ for $x\approx x_0$, in (27) we can set

$$ \begin{equation*} C\leqslant 1, \quad \frac1n\|B\|_1\leqslant 3.05, \quad M\leqslant 3910, \quad M'\leqslant 94.6 \quad\text{and}\quad M_f\leqslant 258. \end{equation*} \notag $$
Then we obtain the error estimate presented in the last column of Table 7. We see that it is reasonably close to the actual error, which is presented in the 4th column.

Table 7

$n$PrecisionTime (sec.)$n$th approximation of $H(x_0)-1$Error estimate
100240.299391$1.44\times 10^{-10}$$6.95\times 10^{-4}$
200366.759046$5.01\times 10^{-22}$$5.60\times 10^{-15}$
3004821.77949$1.73\times 10^{-33}$$3.39\times 10^{-26}$
4006051.22560$6.02\times 10^{-45}$$1.82\times 10^{-37}$
50072115.5499$2.09\times 10^{-56}$$9.19\times 10^{-49}$
60084231.5893$7.26\times 10^{-68}$$4.45\times 10^{-60}$
70096380.6020$2.52\times 10^{-79}$$2.09\times 10^{-71}$
800108608.9937$8.78\times 10^{-91}$$9.65\times 10^{-83}$
900120869.7188$3.06\times 10^{-102}$$4.38\times 10^{-94}$
10001321072.923$1.06\times 10^{-113}$$1.96\times 10^{-105}$
11001441456.021$3.72\times 10^{-125}$$8.70\times 10^{-117}$
12001561852.763$1.29\times 10^{-136}$$3.83\times 10^{-128}$

§ 5. Approximate solutions of Fredholm integral equations with analytic kernels

5.1. Error estimates. Generalities

The notation in this subsection is independent of the notation in the rest of the paper.

Lemma 5.1. Let $f$ be a holomorphic function in a neighbourhood of an annulus $R_1 < |z| < R_2$, and let $f(z)=\sum_{n\in\mathbb{Z}} c_n z^n$ be its Laurent series. Then for $R_1<r<R_2$ and any $n>0$,

$$ \begin{equation} \biggl| c_0-\frac{1}{n}\sum_{k=1}^n f(r\omega^k)\biggr| =\biggl|\int_0^1 f(re^{2\pi it})\,dt-\frac{1}{n}\sum_{k=1}^n f(r\omega^k)\biggr| \leqslant \frac{M_1 q_1^n}{1-q_1^n}+\frac{M_2 q_2^n}{1-q_2^n}, \end{equation} \tag{21} $$
where $\omega=e^{2\pi i/n}$, $q_1=R_1/r$, $q_2=r/R_2$ and $M_j=\max_{|z|=R_j} |f(z)|$ for $j=1,2$.

Proof. We have
$$ \begin{equation*} \sum_{k=1}^n f(r\omega^k)=\sum_{k=1}^n\; \sum_{m\in\mathbb Z} c_m (r\omega^k)^m \quad\text{and}\quad \sum_{k=1}^n \omega^{km}= \begin{cases} n &\text{if $n$ divides $m$}, \\ 0 &\text{otherwise}, \end{cases} \end{equation*} \notag $$
hence the left-hand side of (21) is equal to $\bigl|\sum_{p\in\mathbb Z\setminus\{0\}} c_{pn}r^{pn}\bigr|$ and the coefficients can be estimated using Cauchy integrals.

The lemma is proved.

Lemma 5.2. Let $h>0$, and let $D\subset\mathbb{R}^d$ be a product of segments $[0,n_1h]\times\dots\times[0,n_dh]$ for some positive integers $n_1,\dots,n_d$. Let $f\colon D\to\mathbb{R}$ be a function of class $\mathcal C^2$, and let $M=\max_{i,j}\max_D|\partial_i\partial_j f|$. Then

$$ \begin{equation*} \min_D f \geqslant \min_{h\mathbb Z^d} f-\frac18Md^2h^2, \end{equation*} \notag $$
where $h\mathbb{Z}^d=\{h\vec n\mid\vec n\in\mathbb{Z}^n\}$. A similar estimate holds for $\max_D f$.

Proof. We use induction on $d$. Let the minimum be attained at $x_0\in D$. If $x_0$ lies in the interior of $D$, then we estimate $|f(x)-f(x_0)|$ for the grid point $x$ closest to $x_0$ by using the second-order version of the Taylor-Lagrange formula for $f(x_0+t(x-x_0))$ at $t=0$. If $x_0$ is on the boundary of $D$, then we apply the induction hypothesis to the restriction of $f$ to the facet of $D$ containing $x_0$.

The lemma is proved.

5.2. Error estimates for approximate solutions of Fredholm’s equations

Let $\varphi\colon\mathbb{R}\to\mathbb{C}$ be a continuous solution of Fredholm’s integral equation

$$ \begin{equation} \varphi(x)=\int_0^1 K(x,y)\varphi(y)\,dy+f(x) \end{equation} \tag{22} $$
with analytic complex-valued functions $K$ and $f$ which are (bi)-periodic with period $1$, that is, $K(x,y)=K(x+1,y)=K(x,y+1)$ and $f(x)=f(x+1)$. Assume that $K$ and $f$ extend to complex analytic functions in a neighbourhood of $(D\times\mathbb{R})\cup(\mathbb{R}\times D_1)$ in $\mathbb{C}^2$ and a neighbourhood of $D$ in $\mathbb{C}$, respectively, where
$$ \begin{equation*} D =\{z\in\mathbb C\mid -a \leqslant \operatorname{Im} z\leqslant a \} \quad\!\!\!\text{and}\!\!\!\quad D_1=\{z\in\mathbb C\mid -a_1\leqslant \operatorname{Im} z\leqslant a_1\}, \quad\!\! 0<a_1<a. \end{equation*} \notag $$
Set
$$ \begin{equation*} C=\int_0^1|\varphi(x)|\,dx, \quad M=\max{D\times\mathbb R}|K|, \quad M'_1=\max_{\mathbb R\times D_1}|K| \quad\text{and}\quad M_f=\max_{D}|f|. \end{equation*} \notag $$

Lemma 5.3. The function $\varphi$ extends analytically to a neighbourhood of $D$ and

$$ \begin{equation} M_{\varphi}:=\max_{D_1}|\varphi| \leqslant \frac{a(CM+M_f)}{a-a_1}. \end{equation} \tag{23} $$

Proof. For any $(x_0,y_0)\in\mathbb{R}^2$ and any $n$ we have
$$ \begin{equation*} |\partial_x^n K(x_0,y_0)|\leqslant\biggl|\frac{n!}{2\pi i}\int_{|z|=a}\frac{K(z,y_0)\,dz}{z^{n+1}}\biggr| \leqslant \frac{Mn!}{a^n}, \end{equation*} \notag $$
and similarly $|f^{(n)}(x_0)|\leqslant M_f n!/a^n$. Then differentiating (22) $n$ times with respect to $x$ we obtain
$$ \begin{equation} \bigl|\varphi^{(n)}(x_0)\bigr| =\biggl|\int_0^1 \partial_x^n K(x_0,y)\varphi(y)\,dy+f^{(n)}(x_0)\biggr| \leqslant\frac{(CM+M_f)n!}{a^n}. \end{equation} \tag{24} $$
Hence the Taylor series of $\varphi$ at $x_0$ converges in the disc $|z-x_0|<a$, and for $|z-x_0|\leqslant a_1$ we have
$$ \begin{equation*} |\varphi(z)|=\biggl|\sum_{n\geqslant0}\frac{\varphi^{(n)}(x_0)}{n!}(z-x_0)^n\biggr| \leqslant \sum_{n\geqslant 0}\frac{(CM+M_f)a_1^n}{a^n} =\frac{a(CM+M_f)}{a-a_1}, \end{equation*} \notag $$
which yields the required bound for $M_\varphi$.

The lemma is proved.

Given a positive integer $n$, let us see what happens if we replace the integral in (22) by the $n$th integral sum. Namely, consider the vectors $\varphi^{[n]}=(\varphi_1^{[n]},\dots,\varphi_n^{[n]})$ and $f^{[n]}=(f_1^{[n]},\dots,f_n^{[n]})$, and the $n\times n$ matrix $K^{[n]}=(K_{jk}^{[n]})_{jk}$ defined by

$$ \begin{equation*} \varphi^{[n]}_j=\varphi\biggl(\frac jn\biggr), \qquad f_j^{[n]}=f\biggl(\frac jn\biggr) \quad\text{and}\quad K_{jk}^{[n]}=\frac1n K\biggl(\frac jn,\frac kn\biggr). \end{equation*} \notag $$
Let $\widehat\varphi^{[n]} =\bigl(\widehat\varphi^{[n]}_1,\dots,\widehat\varphi^{[n]}_n\bigr)$ be a solution of the equation
$$ \begin{equation} \widehat\varphi^{[n]}=K^{[n]}\widehat\varphi^{[n]}+f^{[n]}. \end{equation} \tag{25} $$
This equation is a discretization of (22), and it is natural to expect that $\widehat\varphi^{[n]}$ approximates $\varphi$ well. Now, following the approach from [5] we estimate the rate of convergence. Our final purpose is to find a good upper bound for the approximation error
$$ \begin{equation*} E_n:= \biggl|\int_0^1 \varphi(x)\,dx-\frac1n\sum_{j=1}^n \widehat\varphi^{[n]}_j\biggr|. \end{equation*} \notag $$

We define the norms $\|\cdot\|_p$ on $\mathbb{C}^n$, $1\leqslant p\leqslant\infty$, in the usual way. For a square matrix $A=(a_{jk})_{jk}$ with complex entries we set

$$ \begin{equation*} \|A\|_1=\sum_{j,k} |a_{jk}| \quad\text{and}\quad \|A\|_2=\biggl(\sum_{j,k}|a_{jk}|^2\biggr)^{1/2}. \end{equation*} \notag $$

Lemma 5.4. (a) Assume that the matrix $A^{[n]}=I-K^{[n]}$ is invertible and denote its inverse by $B^{[n]}$. Then

$$ \begin{equation} E_n\leqslant \frac{2a(CM+M_f)r_1^n}{(a-a_1)(1-r_1^n)}\biggl(1+\frac1n\|B^{[n]}\|_1M'_1\biggr), \qquad r_1=e^{-2\pi a_1}. \end{equation} \tag{26} $$
If $K$ extends analytically to a neighbourhood of $\mathbb{R}\times D$ and $M'=\max_{\mathbb{R}\times D}K$, then
$$ \begin{equation} E_n\leqslant 4\pi e(CM+M_f)\biggl(1+\frac1n\|B^{[n]}\|_1 M'\biggr)\frac{nar^n}{1-er^n}, \qquad r=e^{-2\pi a}. \end{equation} \tag{27} $$
For $n>\alpha_n$,
$$ \begin{equation} C\leqslant\frac{\|\widehat\varphi^{[n]}\|_1+\alpha_nM_f}{n-\alpha_n}, \quad\textit{where } \alpha_n=\frac{2M'_1a\|B^{[n]}\|_1r_1^n}{(a-a_1)(1-r_1^n)}+\frac{1}{4a}. \end{equation} \tag{28} $$

(b) Assume that $\|K^{[n]}\|_2=M_2<1$. Then the matrix $A^{[n]}$ is invertible and $\dfrac1n\|B^{[n]}\|_1\leqslant \dfrac1{1-M_2}$, which implies, in particular, that $\alpha_n<\alpha_0$ for some constant $\alpha_0=\alpha_0(a,a_1,M,M'_1,M_2,M_f)$, and therefore $C$ can be estimated using (28) for $n>\alpha_0$.

Proof. (a) Let
$$ \begin{equation*} \begin{gathered} \, J=\int_0^1\varphi(x)\,dx, \qquad S=\frac1n\sum_j\varphi\biggl(\frac jn\biggr), \qquad \widehat S=\frac1n\sum_j\widehat\varphi\biggl(\frac jn\biggr), \\ \rho=\varphi^{[n]}-\widehat\varphi^{[n]} \quad\text{and}\quad \sigma=A^{[n]}\rho. \end{gathered} \end{equation*} \notag $$
In this notation, $E_n=|J-\widehat S|$. We have
$$ \begin{equation} \begin{aligned} \, \|\sigma\|_\infty &=\|A^{[n]}\varphi^{[n]}-A^{[n]}\widehat\varphi^{[n]}\|_\infty \stackrel{(25)}{=} \|A^{[n]}\varphi^{[n]}-f^{[n]}\|_\infty \nonumber \\ & =\bigl\|K^{[n]}\varphi^{[n]}-(\varphi^{[n]}-f^{[n]})\bigr\|_\infty. \end{aligned} \end{equation} \tag{29} $$
By (22) we have
$$ \begin{equation*} \varphi^{[n]}_j-f^{[n]}_j=\int_0^1 K\biggl(\frac jn,y\biggr)\varphi(y)\,dy, \end{equation*} \notag $$
and the $j$th component of the vector $K^{[n]}\varphi^{[n]}$ is the $n$th integral sum for this integral. Hence applying Lemma 5.1 to the function $K(j/n,z(\zeta))\varphi(z(\zeta))$ after the change of variable $\zeta=e^{2\pi iz}$, we obtain $\|\sigma\|_\infty\leqslant M'_1C_1$ for $C_1=2M_\varphi r_1^n/(1-r_1^n)$, and then
$$ \begin{equation} \|\rho\|_1=\|B^{[n]}\sigma\|_1\leqslant\|B^{[n]}\|_1\times\|\sigma\|_\infty\leqslant M'_1C_1\|B^{[n]}\|_1. \end{equation} \tag{30} $$
Lemma 5.1, as applied to $\varphi(z(\zeta))$, yields $|J - S| \leqslant C_1$. We also have $|S-\widehat S|\leqslant\frac1n\|\rho\|_1$, hence
$$ \begin{equation*} E_n=|J-\widehat S|\leqslant|J-S|+|S-\widehat S| \leqslant C_1+\frac1n\|\rho\|_1\leqslant C_1+\frac1n M'_1C_1\|B^{[n]}\|_1, \end{equation*} \notag $$
which yields (26) after using (23). Setting $a_1=a-1/(2\pi n)$ (so that $r_1=e^{1/n}r$) and $M'_1<M'$ in (26) we obtain (27).

Let us prove (28). It is easy to check that

$$ \begin{equation*} nC\leqslant\|\varphi^{[n]}\|_1+\frac14\max_{\mathbb R}|\varphi'| \leqslant\|\widehat\varphi^{[n]}\|_1 +\|\rho\|_1+\frac14\max_{\mathbb R}|\varphi'|. \end{equation*} \notag $$

Using estimates (30) and (24) for $\|\rho\|_1$ and $|\varphi'|$, respectively, we obtain

$$ \begin{equation*} nC\leqslant\|\widehat\varphi^{[n]}\|_1 +\frac{2M_1M_\varphi\|B^{[n]}\|_1r_1^n}{1-r_1^n}+\frac{CM+M_f}{4a} \stackrel{(23)}{\leqslant} \|\widehat\varphi^{[n]}\|_1+(CM+M_f)\alpha_n. \end{equation*} \notag $$

(b) Now assume that $\|K^{[n]}\|_2= M_2<1$. Then

$$ \begin{equation*} \|B^{[n]}\|_2=\|(I-K^{[n]})^{-1}\|_2=\|I+K^{[n]}+(K^{[n]})^2+\dotsb\|_2\leqslant \frac{1}{1-M_2}. \end{equation*} \notag $$
By Cauchy’s inequality we also have $\|B^{[n]}\|_1\leqslant n\|B^{[n]}\|_2$.

The lemma is proved.

5.3. A numerical criterion for the existence and uniqueness of solutions

Here we keep the above assumptions about $K(x,y)$ and $f(x)$, except that we do not assume a priori any longer that equation (22) has a continuous solution $\varphi$.

Let $\mathcal K\colon\mathcal C([0,1])\,{\to}\,\mathcal C([0,1])$ be the Fredholm integral operator with kernel $K(x,y)$, that is, the operator $\varphi\mapsto\psi$ where $\displaystyle\psi(x)=\int_0^1 K(x,y)\varphi(y)\,dy$.

Lemma 5.5 (cf. [5], Ch. II, § 1, eq. (26)). Assume that there exists $n$ such that the matrix $I-K^{[n]}$ is invertible and $\alpha_n<n$ where $\alpha_n$ is defined in (28) (note that neither $f$ nor $\varphi$ is used in the definition of $\alpha_n$). Then $1$ is not an eigenvalue of $\mathcal K$, and therefore for any continuous function $f$ equation (22) has a unique continuous solution $\varphi$.

Proof. Let $n$ be such that $\alpha_n<n$. We use Lemma 5.4, (a) for $f=0$, and therefore for $\widehat\varphi^{[n]}=0$. Then (28) reads $C\leqslant 0$, which means that the equation $\mathcal K\varphi=\varphi$ has no nonzero solutions, that is, $1$ is not an eigenvalue of $\mathcal K$. By Fredholm’s theorem (see [2]), in this case (22) has a unique continuous solution for any $f$.

The lemma is proved.

5.4. Analyticity of solutions with respect to a parameter

Let $\Lambda$ be a domain in $\mathbb{C}$, and let $U=\{z\in\mathbb{C}\mid -a\leqslant\operatorname{Im} z\leqslant a\}$, where $a>0$. Let $K(\lambda,x,y)$ be an analytic function in a neighbourhood of $\Lambda\times U^2$ in $\mathbb{C}^3$ and $f(\lambda,y)$ be an analytic function in a neighbourhood of $\Lambda\times U$ in $\mathbb{C}^2$. We assume that $K(\lambda_0,x,y)$ is $(1,1)$-biperiodic and $f(\lambda_0,x)$ is $1$-periodic for any fixed $\lambda_0\in\Lambda$.

For $\lambda\in\Lambda$ let $\mathcal K_\lambda\colon\mathcal C([0,1])\to\mathcal C([0,1])$ be the Fredholm integral operator ${\varphi\mapsto\psi}$, where $\displaystyle\psi(x)=\int_0^1 K(\lambda,x,y)\varphi(y)\,dy$. The next lemma follows directly from Fredholm’s results in his seminal paper [2] (a more general fact was proved in [11]).

Lemma 5.6. Suppose that $1$ is not an eigenvalue of $\mathcal K_\lambda$ for any $\lambda\in\Lambda$. Then for any $\lambda\in\Lambda$ there exists a unique solution $\varphi(\lambda,x)$ of the equation

$$ \begin{equation} \varphi(\lambda,x)=\int_0^1 K(\lambda,x,y)\varphi(\lambda,y)\,dy+f(\lambda,x) \end{equation} \tag{31} $$
and the function $\varphi(\lambda,x)$ is analytic in a neighbourhood of $\Lambda\times U$.

Proof. By Fredholm’s results (see [2] and also [6]), for any $\lambda\in\Lambda$ the solution $\varphi(\lambda,t)$ is unique under our assumptions, and it can be expressed as
$$ \begin{equation*} \varphi(\lambda,x)=f(\lambda,x)+\int_0^1\frac{D(\lambda,x,y)}{D(\lambda)}f(\lambda,y)\,dy, \end{equation*} \notag $$
where
$$ \begin{equation} \begin{gathered} \, D(\lambda)=\sum_{n=0}^\infty\frac{(-1)^n A_n(\lambda)}{n!}, \qquad D(\lambda,x,y)=\sum_{n=0}^\infty\frac{(-1)^n B_n(\lambda,x,y)}{n!}, \\ \notag A_n(\lambda)=\int_{[0,1]^n} K(\lambda,\boldsymbol x,\boldsymbol x)\,d\boldsymbol x, \qquad B_n(\lambda,x,y)=\int_{[0,1]^n} K(\lambda,x,\boldsymbol x,y,\boldsymbol x)\,d\boldsymbol x \end{gathered} \end{equation} \tag{32} $$
and
$$ \begin{equation*} K(\lambda,x_1,\dots,x_n,y_1,\dots,y_n)=\det\bigl(K(\lambda,x_i,y_j)\bigr)_{i,j=1}^n. \end{equation*} \notag $$
It was shown in [2] that $D(\lambda)$ does not vanish on $U$ (because $1$ is not an eigenvalue of $\mathcal K_\lambda$ for any $\lambda\in\Lambda$). It is clear that the functions $A_n$ and $B_n$ are analytic in $\Lambda$ and $\Lambda\times U^2$, respectively, and Hadamard’s inequality $|\det N|\leqslant n^{n/2}\max_{i,j}|N_{ij}|$ implies the upper bounds
$$ \begin{equation*} |A_n(\lambda)|\leqslant n^{n/2}M(\lambda)^n \quad\text{and}\quad |B_{n-1}(\lambda,x,y)|\leqslant n^{n/2}M(\lambda)^n \end{equation*} \notag $$
(cf. [2], p. 368, line 4), where $M(\lambda)=\sup_{(x,y)\in U^2}|K(\lambda,x,y)|$. Hence the series in (32) converge to analytic functions, which yields the result.

The lemma is proved.

§ 6. Nonprimitive lattice triangulations

Denote the number of all (not necessarily primitive) lattice triangulations of the $m\times n$ rectangle by $f^{\mathrm{np}}(m,n)$ and set

$$ \begin{equation*} c^{\mathrm{np}}=\lim_{n\to\infty}\frac{\log_2 f^{\mathrm{np}}(n,n)}{n^2}. \end{equation*} \notag $$

Proposition 6.1. The inequality $c^{\mathrm{np}} \leqslant 4.735820221\dots$ holds.

Proof. Let $N=n^2$. Any lattice triangulation can be subdivided to a primitive lattice triangulation. Hence a lattice triangulation is completely determined by a choice of a primitive lattice triangulation and a set of edges to be removed. Let $f_k^{\mathrm{np}}(n,n)$ be the number of lattice triangulations of an $n\times n$ square that have $k$ interior vertices and therefore $\approx 3k$ edges. Then
$$ \begin{equation} f_k^{\mathrm{np}}(n,n) \leqslant 2^{cN} \binom{3N}{3k} \end{equation} \tag{33} $$
(recall that $2^{cN}$ is a bound for the number of primitive lattice triangulations). On the other hand the number of triangulations with vertices in an arbitrary fixed set of $k$ points on a plane is bounded above by $30^k$ (see [10]), hence
$$ \begin{equation} f_k^{\mathrm{np}}(n,n) \leqslant 30^k \binom{N}{k}. \end{equation} \tag{34} $$
Combining (33) and (34) with Stirling’s formula we obtain
$$ \begin{equation} \begin{gathered} \, c^{\mathrm{np}} \leqslant \max_{0\leqslant x\leqslant 1} \min\bigl(3h(x)+c,h(x)+x\log_2 30\bigr), \\ h(x)=-x\log_2 x-(1-x)\log_2(1-x). \notag \end{gathered} \end{equation} \tag{35} $$
Using the bound $c\leqslant 4\log_2 \frac{1+\sqrt5}2$ (see [7], [12] and [13]) we obtain the required result (the maximum in (35) is attained at $x=0.83206855$).

The proposition is proved.


Bibliography

1. E. E. Anclin, “An upper bound for the number of planar lattice triangulations”, J. Combin. Theory Ser. A, 103:2 (2003), 383–386  crossref  mathscinet  zmath
2. I. Fredholm, “Sur une classe d'équations fonctionnelles”, Acta Math., 27:1 (1903), 365–390  crossref  mathscinet  zmath
3. I. M. Gelfand, M. M. Kapranov and A. V. Zelevinsky, Discriminants, resultants, and multidimensional determinants, Math. Theory Appl., Birkhäuser Boston, Inc., 1994, x+523 pp.  crossref  mathscinet  zmath
4. V. Kaibel and G. M. Ziegler, “Counting lattice triangulations”, Surveys in combinatorics 2003 (Univ. of Wales, Bangor, UK 2003), London Math. Soc. Lecture Note Ser., 307, Cambridge Univ. Press, Cambridge, 2003, 277–307  crossref  mathscinet  zmath
5. L. V. Kantorovich and V. I. Krylov, Approximate methods of higher analysis, 5th corr. ed., Fizmatgiz, Moscow–Leningrad, 1962, 708 pp.  mathscinet  zmath; English transl. of 3rd ed., Interscience Publishers, Inc., New York; P. Noordhoff Ltd., Groningen, 1958, xv+681 pp.  mathscinet  zmath
6. B. V. Khvedelidze, “Fredholm equation”, Matematicheskaya Entsiklopediya, ed. I. M. Vinogradov, Sov. Entsiklopediya, Moscow, 1977; English transl., Fredholm equation, Encyclopedia of Mathematics http://encyclopediaofmath.org/index.php?title=Fredholm_equation&oldid=46977
7. J. Matoušek, P. Valtr and E. Welzl, On two encodings of lattice triangulations, Manuscript, 2006
8. S. Yu. Orevkov, “Asymptotic number of triangulations with vertices in $\mathbf Z^2$”, J. Combin. Theory Ser. A, 86:1 (1999), 200–203  crossref  mathscinet  zmath
9. S. Yu. Orevkov and V. M. Kharlamov, “Asymptotic growth of the number of classes of real plane algebraic curves as the degree grows”, Representation theory, dynamical systems, combinatorial and algorithmic methods. Part V, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 266, St. Petersburg Department of Steklov Mathematical Institute, St. Petersburg, 2000, 218–233  mathnet  mathscinet  zmath; English transl. in J. Math. Sci. (N.Y.), 113:5 (2003), 666–674  crossref
10. M. Sharir and A. Sheffer, “Counting triangulations of planar point sets”, Electron. J. Combin., 18:1 (2011), 70, 74 pp.  crossref  mathscinet  zmath
11. J. D. Tamarkin, “On Fredholm's integral equations, whose kernels are analytic in a parameter”, Ann. of Math. (2), 28:1–2 (1926–1927), 127–152  crossref  mathscinet  zmath
12. E. Welzl, “The number of triangulations on planar point sets”, Graph drawing, Lecture Notes in Comput. Sci., 4372, Springer, Berlin, 2007, 1–4  crossref  mathscinet  zmath
13. E. Welzl, J. Matušek and P. Valtr, Lattice triangulations, Talk in Freie Univ. (Berlin 2006)

Citation: S. Yu. Orevkov, “Counting lattice triangulations: Fredholm equations in combinatorics”, Mat. Sb., 213:11 (2022), 50–78; Sb. Math., 213:11 (2022), 1530–1558
Citation in format AMSBIB
\Bibitem{Ore22}
\by S.~Yu.~Orevkov
\paper Counting lattice triangulations: Fredholm equations in combinatorics
\jour Mat. Sb.
\yr 2022
\vol 213
\issue 11
\pages 50--78
\mathnet{http://mi.mathnet.ru/sm9727}
\crossref{https://doi.org/10.4213/sm9727}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4582605}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022SbMat.213.1530O}
\transl
\jour Sb. Math.
\yr 2022
\vol 213
\issue 11
\pages 1530--1558
\crossref{https://doi.org/10.4213/sm9727e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992276000004}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165921888}
Linking options:
  • https://www.mathnet.ru/eng/sm9727
  • https://doi.org/10.4213/sm9727e
  • https://www.mathnet.ru/eng/sm/v213/i11/p50
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:271
    Russian version PDF:30
    English version PDF:50
    Russian version HTML:150
    English version HTML:81
    References:32
    First page:10
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024