To verify our ansatz, we will compute the Kronecker’s mass and the Robin’s mass for different \(\Omega \). We will compare our analytic results with numerical simulations in Sect. 6. We will start considering flat manifolds having \(g(x)=\mathbb I\), and consider manifolds with uniform curvature starting from Sect. 5.5.
The Unit Rectangle
Let us start by considering the problem on the rectangle. We call \(\mathcal R(\rho )\) the rectangle \([0,\sqrt{\rho }] \times [0,1/\sqrt{\rho }]\), and we consider the Laplace–Beltrami operator with Neumann boundary conditions. The eigenfunctions of \(-\Delta \) on \(\mathcal R(\rho )\) are given by
$$\begin{aligned} u_{(n,m)}(x,y) = \cos \left( \frac{\pi n x}{\sqrt{\rho }} \right) \cos \left( \sqrt{\rho }\pi m y\right) ,\quad (x,y)\in \mathcal R(\rho ),\quad (n,m)\in \mathbb {N}^2\setminus (0,0). \end{aligned}$$
(67)
The corresponding eigenvalues are
$$\begin{aligned} \lambda _{(n,m)} = \pi ^2 \left( \rho m^2 + \frac{n^2}{\rho } \right) ,\quad (n,m)\in \mathbb {N}^2\setminus (0,0). \end{aligned}$$
(68)
We proceed computing the Kronecker mass using the regularized function
$$\begin{aligned} \begin{aligned} Z(s)&=\left( \frac{\rho }{\pi ^2}\right) ^s\sum _{\begin{array}{c} (n,m)\in \mathbb {N}^2\\ n^2+m^2\ne 0 \end{array}}\frac{1}{(\rho ^2m^2 + n^2)^s}\\&=\frac{1}{4}\left( \frac{\rho }{\pi ^2}\right) ^s\sum _{\begin{array}{c} (n,m)\in \mathbb {Z}^2\\ n^2+m^2\ne 0 \end{array}}\frac{1}{(\rho ^2m^2 + n^2)^s} +\frac{1}{2}\left( \frac{\rho }{\pi ^2}\right) ^s\left( \sum _{m \ge 1}\frac{1}{(\rho ^2m^2)^s}+\sum _{n \ge 1}\frac{1}{(n^2)^s}\right) \\&=\frac{\zeta _\tau (s)}{4\pi ^{2s}}+\frac{\rho ^s+\rho ^{-s}}{2\pi ^{2s}}\sum _{n=1}^\infty \frac{1}{n^{2s}}. \end{aligned} \end{aligned}$$
(69)
Here we have adopted \(\tau =i\rho \), in compliance with standard notation for modular forms, and have introduced the lattice zeta function \(\zeta _\tau (s)\) defined in Appendix D. This calculation is readily performed thanks to a remarkable result due to Kronecker (and known as first limit formula of Kronecker), reported in Appendix D, Eq. (D.7), and that we repeat here:
$$\begin{aligned} \zeta _\tau (s):=\sum _{\begin{array}{c} (n,m)\in \mathbb {Z}^2\\ n^2+m^2\ne 0 \end{array}}\frac{[\mathfrak {I}(\tau )]^s}{|n + \tau m|^{2s}} =\frac{\pi }{s-1}+2 \pi \left[ \gamma _{\text {E}} - \ln ( 2\sqrt{\mathfrak {I}(\tau )}|\eta (\tau )|^2 )\right] + o(s-1), \end{aligned}$$
(70)
where \(\eta (\tau )\) is the Dedekind \(\eta \) function. Kronecker’s formula allows us to immediately obtainFootnote 8
$$\begin{aligned} K_{\mathcal R}(\rho )=\frac{\gamma _{\text {E}}}{2 \pi } - \frac{\ln (4\pi ^2\rho |\eta (i \rho )|^4)}{4\pi } + \frac{1}{12} \left( \rho + \frac{1}{\rho } \right) \end{aligned}$$
(71)
that for \(\rho =1\) (unit square) simplifies to
$$\begin{aligned} K_{\mathcal R}:=K_{\mathcal R}(1) =\frac{\gamma _{\text {E}}}{2\pi }+\frac{\ln (4\pi )}{4\pi }-\frac{\ln \Gamma \left( 1\big /4\right) }{\pi } + \frac{1}{6}. \end{aligned}$$
(72)
We will see in the following that the first limit formula of Kronecker will allow us to extract the Kronecker’s mass for many types of flat domains: this explains our choice of ‘Kronecker mass’ for denoting \(K_\Omega \).
We can give a slightly more compact form to the function in (71), shifted by its minimal value above:
$$\begin{aligned} K_{\mathcal R}(\rho ) - K_{\mathcal R}(1) = -\frac{1}{2\pi } \ln \frac{\eta (i \rho ) \eta \left( i\rho ^{-1} \right) }{\eta ^2(i)}+ \frac{1}{12}\left( \sqrt{\rho } - \frac{1}{\sqrt{\rho }}\right) ^2. \end{aligned}$$
(73)
We shall make a remark on this expression. In the limit in which the rectangle is very elongated, we getFootnote 9
$$\begin{aligned} \lim _{\rho \rightarrow \infty }\frac{2K_{\mathcal R}(\rho )}{\rho }=\lim _{\rho \rightarrow 0} 2 \rho K_{\mathcal R}(\rho )= \frac{1}{3}, \end{aligned}$$
(74)
that is the average cost for the Poisson–Poisson one-dimensional assignment problem on the segment of unit length [39]. This is not by accident. Indeed, in any rectangular domain we can evaluate the average energy of the permutation in which the k-th red point counting from the left is matched to the k-th blue point counting from the left. This configuration is optimal w.h.p. in the limit \(\rho \rightarrow +\infty \), and would be optimal, at any \(\rho \), if the vertical coordinates of all the points were equal. On the other side, a worst case is when all the vertical coordinates of red points are zero, and all the vertical coordinates of blue points are \(1/\sqrt{\rho }\), so that, calling \(E_{[0,1]}(N)\) the average energy for the 1-dimensional problem on the [0, 1] segment, we get
$$\begin{aligned} \rho E_{[0,1]}(N)\le E_{{\mathcal R}(\rho )}(N) \le \rho E_{[0,1]}(N) + \frac{N}{\rho } \end{aligned}$$
(75)
which, by substituting our scaling ansatz, gives
$$\begin{aligned} \rho E_{[0,1]}(N)\le \frac{1}{2 \pi } \ln N + 2c_*(N) + 2K_{\mathcal R}(\rho )\le \rho E_{[0,1]}(N) + \frac{N}{\rho }. \end{aligned}$$
(76)
When we take a limit \(N \rightarrow \infty \), \(\rho \rightarrow \infty \) on a direction \(\rho \gg \sqrt{N}\), we thus get
$$\begin{aligned} \frac{1}{3} \le \frac{2K_{\mathcal R}(\rho )}{\rho } + \mathcal {O}\left( \frac{\ln N}{\rho } \right) \le \frac{1}{3} +\mathcal {O}\left( \frac{N}{\rho ^2} \right) , \end{aligned}$$
(77)
which is consistent with (74). In the following we will encounter various other domains which allow a consistency check with a 1-dimensional limit. We will reach similar conclusions, without entering in the details of the estimates, as this is done by minor modifications of the reasonings presented here.
The Flat Torus
We shall now consider the problem on the flat torus \(\mathcal T(\tau )\). To describe the corresponding manifold, let us first consider the lattice of points on \(\mathbb {R}^2\)
$$\begin{aligned} \Lambda = \left\{ \underline{\omega }\ n,\quad n\in \mathbb {Z}^2\right\} \end{aligned}$$
(78)
generated by the matrix
corresponding to the base vectors
In such lattice it is possible to define fundamental parallelograms
, containing no further lattice points in its interior or boundary. A fundamental parallelogram is given for example by
so that
, see Fig. 2a. We will also use a shortcut adapted to rectangles,
A torus \(\mathcal T\) is defined as a quotient between the complex plane and a lattice \(\Lambda \), \(\mathcal T:=\mathbb {R}^2/\Lambda \). In other words, each point
is identified with the set of points \(\{x+\underline{\omega }\ n,\ n\in \mathbb {Z}^2\}\), the distance between two points in
being the minimum distance between the elements of their equivalence classes. It is well known that two matrices \(\underline{\omega }\) and \(\underline{\omega }'\) identify the same lattice \(\Lambda \) and the same torus \(\mathcal T\) (although not the same fundamental domain
) if and only if \((\underline{\omega })^{-1} \underline{\omega }' \in \text {SL}(2,\mathbb {Z})\). For each \(\underline{\omega }\), we introduce the half-period ratio
$$\begin{aligned} \tau :=\frac{s+ih}{\ell }\in \mathbb {C}. \end{aligned}$$
(83)
Given a lattice \(\Lambda \) generated by \(\underline{\omega }\), it is possible to associate to it a dual lattice \(\Lambda ^*\) generated by \(\underline{\omega }^*\), such that \(\underline{\omega }^*\ \underline{\omega }^T=\mathbb I\), identity matrix, i.e.,
Each torus \(\mathcal T=\mathbb {R}^2/\Lambda \) is then naturally associated to a dual torus given by \(\mathcal T^*:=\mathbb {R}^2/\Lambda ^*\).
In the following, we will restrict, without loss of generality, to the case in which the fundamental parallelograms have unit area, choosing
such that \(\rho \in \mathbb {R}^+\) and \(\sigma \in \mathbb {R}\), and we will denote the corresponding torus by \(\mathcal T(\tau )\), where \(\tau :=\sigma +i\rho \) is the half-period ratio.
The Kronecker’s mass. Due to the periodicity conditions, the eigenfunctions of \(-\Delta \) on \(\mathcal T(\tau )\) have the form
$$\begin{aligned} u_{k^*}(x) = \exp (2 \pi i\; k^*\cdot x) \end{aligned}$$
(86)
for all \(k^*=\underline{\omega }^*\ k\in \Lambda ^*\), \(k=\left( {\begin{array}{c}n\\ -m\end{array}}\right) \in \mathbb {Z}^2\). The corresponding eigenvalue is
$$\begin{aligned} \lambda _{(n,m)}=|2 \pi k^*|^2 =(2\pi )^2 \frac{|n+\tau m|^2}{\rho } =(2\pi )^2 \frac{|n+\tau m|^2}{\mathfrak {I}(\tau )}. \end{aligned}$$
(87)
We can compute now the Kronecker mass using the regularized function
$$\begin{aligned} Z(s)=\sum _{k^*}\frac{1}{|2\pi k^*|^{2s}}=\frac{1}{(2\pi )^{2s}}\sum _{\begin{array}{c} (m,n)\in \mathbb {Z}^2\\ n^2+m^2\ne 0 \end{array}}\frac{[\mathfrak {I}(\tau )]^s}{|n+\tau m|^{2s}} \end{aligned}$$
(88)
and removing the pole in \(s\rightarrow 1\), as discussed in Sect. 3. This calculation is readily performed, again thanks to the first limit formula of Kronecker, Eq. (D.7), which allows us to immediately obtain
$$\begin{aligned} K_{\mathcal T}(\tau ):=\frac{\gamma _{\text {E}}}{2 \pi } -\frac{1}{4\pi } \ln \left( 16\pi ^2\mathfrak {I}(\tau ) |\eta (\tau )|^4 \right) . \end{aligned}$$
(89)
In Fig. 2b we present a contour-plot of the related expression \(\mathfrak {I}(\tau ) |\eta (\tau )|^4\) in the complex plane \(\tau \), confined to the canonical fundamental region of the moduli space. In particular, this function diverges for \(\tau \rightarrow 0\) and has minimum at \(\tau =\exp (i \pi (\frac{1}{2} \pm \frac{1}{6}))\). This implies that, among all unit tori equipped with the flat metric, the “hexagonal” one, that is the one for which \(\tau \) is a sixth root of unity, is the one in which the average cost of the Euclidean Random Assignment Problem is minimised. More strikingly, as deduced from results in [5] which are in turn based on the results in [35], the hexagonal torus is minimal also among unit surfaces with non-uniform metric.
Example: the rectangular and rhomboidal tori. We shall call “rectangular torus” a torus in which the fundamental parallelogram is a rectangle. This case corresponds to \(\tau =i\rho \), with \(\rho >0\) real. Our formula specialises to
$$\begin{aligned} K_{\mathcal T}(i\rho )=\frac{\gamma _{\text {E}}-\ln (4\pi \sqrt{\rho })}{2\pi } - \frac{1}{\pi } \ln |\eta (i \rho )|, \end{aligned}$$
(90)
which is invariant under the map \(\rho \mapsto \rho ^{-1}\), as it should. In the region \(\rho \in (0,1]\) the lowest value is achieved at \(\rho =1\) (see also Fig. 2b), where
$$\begin{aligned} K_{\mathcal T}:=K_{\mathcal T}(i)=\frac{\gamma _{\text {E}}}{2\pi }+ \frac{\ln \pi }{4 \pi } - \frac{1}{\pi } \ln \Gamma \left( 1\big /4\right) . \end{aligned}$$
(91)
We can give a slightly more compact form to this function, shifted by its minimal value:
$$\begin{aligned} K_{\mathcal T}(i\rho ) - K_{\mathcal T}(i) = -\frac{\ln \rho }{4\pi } -\frac{1}{\pi } \ln \frac{\eta (i \rho ) }{\eta (i)}= -\frac{1}{2\pi } \ln \frac{\eta (i \rho ) \eta \left( i\rho ^{-1} \right) }{\eta ^2(i)}. \end{aligned}$$
(92)
Similarly, we shall call “rhomboidal torus” a torus in which the fundamental parallelogram is a rhombus. This case corresponds to \(\tau =\hbox {e}^{i \theta }\), with \(0< \theta \le \pi \big /2\), and our formula specialises to
$$\begin{aligned} K_{\mathcal T}(\hbox {e}^{i\theta })= \frac{\gamma _{\text {E}}-\ln (4\pi )}{2\pi }- \frac{1}{4\pi } \ln \sin \theta - \frac{1}{\pi } \ln |\eta (\hbox {e}^{i \theta })|, \end{aligned}$$
(93)
that is, again shifting by the value for the standard torus,
$$\begin{aligned} K_{\mathcal T}(\hbox {e}^{i\theta })-K_{\mathcal T}(i) =- \frac{1}{4\pi } \ln \sin \theta - \frac{1}{\pi } \ln \frac{2 \pi ^{3\big /4}|\eta (\hbox {e}^{i \theta })|}{\Gamma (1\big /4)}. \end{aligned}$$
(94)
As was the case for the rectangle, the expression in Eq. (92), in the limit in which the torus is very “thin and long”, becomes
$$\begin{aligned} \lim _{\rho \rightarrow \infty }\frac{2K_{\mathcal T}(i\rho ) }{\rho } =\lim _{\rho \rightarrow 0} 2 \rho K_{\mathcal T}(i\rho ) = \frac{1}{6}. \end{aligned}$$
(95)
This happens to be the average cost for the Poisson–Poisson one-dimensional assignment problem on the circle of unit length [39], as was to be expected, by a reasoning analogous to the one presented for the case of the rectangle.
The Robin mass. Let us now evaluate, for the generic flat torus \(\mathcal T(\tau )\), the Robin mass \(R_{\mathcal T}(\tau )\). Calling \(z=z(x,y)= (x_1 -y_1) + i(x_2 - y_2)\), the Green’s function on the torus is given in this case by [40]
$$\begin{aligned} G(x,y) =-\frac{1}{2\pi } \ln \frac{\theta _1\left( \sqrt{\mathfrak {I}(\tau )} \, z ; \tau \right) }{\eta (\tau )} +\frac{\mathfrak {I}(\tau ) \, (\mathfrak {I}(z))^2}{2} \end{aligned}$$
(96)
where \(\theta _1(z; \tau )\) is an elliptic \(\theta \) function. The Robin mass is obtained from
$$\begin{aligned} R_{\mathcal T}(\tau ):=-\lim _{z\rightarrow 0} \left[ \frac{1}{2\pi } \ln \left| \frac{\theta _1\left( \sqrt{\mathfrak {I}(\tau )} z; \tau \right) }{\eta (\tau )} \right| -\frac{\ln |z| }{2\pi } \right] = -\frac{1}{4\pi } \ln \left[ 4\pi ^2\mathfrak {I}(\tau )|\eta (\tau )|^4\right] . \end{aligned}$$
(97)
It is immediately seen that, in agreement with the Morpurgo theorem, Eq. (55) is satisfied.
Other Boundary Conditions on the Unit Rectangle
The unit rectangle and the rectangular torus are obtained starting from the fundamental domain
in Eq. (82), and assuming respectively open (i.e., Neumann for the field \(\phi \)) and periodic boundary conditions. Other choices of boundary conditions are possible, which correspond to other classical surfaces, with or without boundary. Each choice leads to a different spectrum of the Laplacian, which in turn implies a different Kronecker mass (and, according to our theory, a different finite-size correction to the optimal cost of the assignment problem).
The Cylinder
Let us consider the domain
and let us take periodic boundary conditions in the horizontal direction (i.e., the side of length \(\sqrt{\rho }\)) and Neumann boundary conditions in the vertical direction (i.e., the side of length \(1\big /\sqrt{\rho }\)), see Fig. 4a. The resulting surface is a cylinder, that we shall denote by \(\mathcal C(\rho )\). The eigenfunctions of \(-\Delta \) are the set of functions
$$\begin{aligned} u_{(m,n)}(x,y) =\exp \left( \frac{2i \pi m x}{\sqrt{\rho }}\right) \cos \left( \pi \sqrt{\rho }n y\right) ,\qquad m\in \mathbb {Z},\ n\in \mathbb {N}. \end{aligned}$$
(98)
The corresponding eigenvalues are therefore
$$\begin{aligned} \lambda _{(m,n)} = \pi ^2 \left( 4 \frac{m^2}{\rho } + \rho n^2 \right) , \qquad m\in \mathbb {Z},\ n\in \mathbb {N}. \end{aligned}$$
(99)
Repeating the same type of calculations performed for the rectangle (that is, expressing the regularised sum as a combination of \(\zeta _{\tau }(s)\) (for some \(\tau \)’s) and \(\zeta (2s)\)), we obtain
$$\begin{aligned} K_{\mathcal C}(\rho ) =\frac{\gamma _{\text {E}}}{2 \pi } - \frac{\ln (16 \pi ^2 \rho )}{4\pi } - \frac{1}{\pi } \ln \eta (2i \rho ) + \frac{1}{24\rho } \end{aligned}$$
(100)
so that
$$\begin{aligned} K_{\mathcal C}:=K_{\mathcal C}(1)=\frac{\gamma _{\text {E}}}{2\pi } + \frac{3 \ln 2}{8 \pi } + \frac{\ln \pi }{4 \pi } - \frac{\ln \Gamma \left( 1\big /4 \right) }{\pi } + \frac{1}{24}. \end{aligned}$$
(101)
We also remark that
$$\begin{aligned} \lim _{\rho \rightarrow \infty } \frac{2K_{\mathcal C}(\rho )}{\rho } = \frac{1}{3}, \end{aligned}$$
(102)
which is the cost density for the one-dimensional assignment problem with open boundary conditions (i.e., on the unit segment), while
$$\begin{aligned} \lim _{\rho \rightarrow 0}2 \rho K_{\mathcal C}(\rho )= \frac{1}{6}, \end{aligned}$$
(103)
which is the density of cost for the one-dimensional assignment problem with periodic boundary conditions (i.e., on the unit circle), again, as was to be expected. The nontrivial solution of Eq. \(K_{\mathcal C}(\rho )=K_{\mathcal C}\) is \(\rho =0.625352\dots \), while the minimum value of the mass occurs for \(\rho =0.793439\dots \)
Remark that the constants above do not appear in the study of [35], because in our context, in presence of a boundary, we should impose Neumann boundary conditions (while the authors of [35] only analyse the case of Dirichlet boundary conditions).
The Möbius Strip
Starting again from the rectangle
, we can identify each point
with all its images in \(\mathbb {R} \times [0,1\big /\sqrt{\rho }]\) generated by the map \((x,y)\rightarrow (x+\sqrt{\rho },1\big /\sqrt{\rho }-y)\). That is, if we see the surface as the fundamental rectangular domain
, we impose open boundary conditions along the horizontal direction, and identify the two vertical sides after a ‘twist’ (that is, the top part on the left is glued to the bottom part on the right). The obtained domain \(\mathcal M(\rho )\) is called the Möbius strip, see Fig. 4b. The eigenfunctions of \(-\Delta \) are
$$\begin{aligned} u_{(m,n)}(x,y) =\exp \left( \frac{i \pi m x}{\sqrt{\rho }}\right) \cos \left( \pi \sqrt{\rho }n y\right) ,\qquad m\in \mathbb {Z},\ n\in \mathbb {N},\ m+n \text {~even.} \end{aligned}$$
(104)
The parity constraint implements the twisted identification of the strip. The corresponding eigenvalues are
$$\begin{aligned} \lambda _{(m,n)} = \pi ^2 \left( \rho ^{-1}m^2 + \rho n^2\right) . \end{aligned}$$
(105)
Repeating now the usual arguments we get
$$\begin{aligned} K_{\mathcal M}(\rho )=\frac{\gamma _{\text{ E }}}{2 \pi } - \frac{\ln (4 \pi ^2 \rho )}{4\pi }- \frac{1}{\pi }\ln \frac{\eta ^3(i \rho )}{\eta (2i \rho )\eta \left( i\frac{\rho }{2}\right) }+ \frac{1}{24 \rho }, \end{aligned}$$
(106)
so that
$$\begin{aligned} K_{\mathcal M}:=K_{\mathcal M}(1)= \frac{\gamma _{\text{ E }}}{2\pi } + \frac{\ln (2\pi )}{4 \pi } - \frac{\ln \Gamma \left( 1\big /4\right) }{\pi } + \frac{1}{24}. \end{aligned}$$
(107)
We also remark that
$$\begin{aligned} \lim _{\rho \rightarrow \infty } \frac{2K_{\mathcal M}(\rho )}{\rho } = \frac{1}{12} \end{aligned}$$
(108)
which is the average cost for the problem on the segment of length \(\frac{1}{2}\) (and the fact that the length is not 1 is related to the fact that the twisted boundary conditions are effectively ‘folding in two’ the segment), while
$$\begin{aligned} \lim _{\rho \rightarrow 0} 2 \rho K_{\mathcal M}(\rho )= \frac{1}{6}, \end{aligned}$$
(109)
which is the cost for the problem on the unit circle. The non-trivial solution of Eq. \(K_{\mathcal M}(\rho ) = K_{\mathcal M}\) is found for \(\rho =4.1861\dots \), whereas the minimum is achieved at \( \rho =2.30422\dots \)
The Klein Bottle
If we identify both pairs of opposite sides of the rectangle, one pair (say, the horizontal sides) in the ordinary way, and the other pair in the twisted way as in the Möbius strip, we obtain the Klein bottle \(\mathcal K(\rho )\), see Fig. 4c. The eigenfunctions of \(-\Delta \) are in this case
$$\begin{aligned} u_{(m,n)}(x,y)=\hbox {e}^{\frac{\pi i m}{\sqrt{\rho }} x}\cos \left( 2 \pi n\sqrt{\rho }y\right) , \qquad m\in \mathbb {Z},\ n\in \mathbb {N},\ m+n\text {\ even} \end{aligned}$$
(110)
and
$$\begin{aligned} v_{(m,n)}(x,y)=\hbox {e}^{\frac{2m+1}{\sqrt{\rho }}\pi i x} \sin \left( 2 \pi n\sqrt{\rho }y\right) ,\qquad m\in \mathbb {Z},\ n\in \mathbb {N}^+. \end{aligned}$$
(111)
Proceeding as above, one can obtain
$$\begin{aligned} K_{\mathcal K}(\rho )=\frac{\gamma _{\text{ E }}}{2 \pi } - \frac{\ln (4 \pi ^2 \rho )}{4\pi } - \frac{1}{\pi }\ln \eta \left( i \frac{\rho }{2}\right) - \frac{\zeta (2)}{2 \pi ^2 \rho } \end{aligned}$$
(112)
so that, in particular,
$$\begin{aligned} K_{\mathcal K}:=K_{\mathcal K}(1)=\frac{\gamma _E}{2 \pi } +\frac{7}{8 \pi } \ln 2 + \frac{1}{4 \pi } \ln \pi - \frac{\ln \Gamma \left( 1\big /4 \right) }{\pi } - \frac{1}{12}. \end{aligned}$$
(113)
We also remark that both one-dimensional limits coincide with the corresponding constructions for the Möbius strip, and indeed the limits of the analytical expressions are the same, as
$$\begin{aligned} \lim _{\rho \rightarrow \infty } \frac{2K_{\mathcal K}(\rho )}{\rho } = \frac{1}{12}, \end{aligned}$$
(114)
while
$$\begin{aligned} \lim _{\rho \rightarrow 0} 2 \rho K_{\mathcal K}(\rho )= \frac{1}{6}. \end{aligned}$$
(115)
Here \(K_{\mathcal K}(\rho ) = K_{\mathcal K}\) for \(\rho = 1.09673\dots \), whereas the minimum is obtained at \(\rho = 1.04689\dots \).
The Boy Surface
As a final example, let us take twisted boundary conditions for both pairs of opposite sides of
. In this way we obtain the so-called Boy surface \(\mathcal B(\rho )\), see Fig. 4d. The eigenfunctions of \(-\Delta \) are
$$\begin{aligned} u_{(m,n)}(x,y)= & {} \cos \left( \frac{ \pi m}{\sqrt{\rho }} x\right) \cos \left( \pi n\sqrt{\rho }y\right) , \qquad m,n\in \mathbb {N},\ m+n\text {\ even,} \end{aligned}$$
(116a)
$$\begin{aligned} v_{(m,n)}(x,y)= & {} \sin \left( \frac{ \pi m}{\sqrt{\rho }} x\right) \cos \left( \pi n\sqrt{\rho }y\right) , \qquad m,n\in \mathbb {N},\ m+n\text {\ odd.} \end{aligned}$$
(116b)
The calculation proceeds as in the other cases, giving
$$\begin{aligned} K_{\mathcal B}(\rho )=\frac{\gamma _{\text{ E }}}{2 \pi } - \frac{\ln (4 \pi ^2 \rho )}{4\pi } - \frac{\ln \eta \left( i \rho \right) }{\pi } -\frac{1}{24}\left( \rho +\frac{1}{\rho } \right) \end{aligned}$$
(117)
which is symmetric for \(\rho \leftrightarrow \frac{1}{\rho }\), as it must be. In particular
$$\begin{aligned} K_{\mathcal B}:=K_{\mathcal B}(1)=\frac{\gamma _{\text{ E }}}{2 \pi } + \frac{3}{8 \pi } \ln 2 + \frac{1}{4 \pi } \ln \pi - \frac{\ln \Gamma \left( 1\big /4 \right) }{\pi } - \frac{1}{12}. \end{aligned}$$
(118)
Now, both one-dimensional limits produce a domain corresponding to the segment of length \(\frac{1}{2}\), and indeed
$$\begin{aligned} \lim _{\rho \rightarrow \infty }\frac{K_{\mathcal B}(\rho )}{\rho }=\lim _{\rho \rightarrow 0} 2 \rho K_{\mathcal B}(\rho ) = \frac{1}{12}. \end{aligned}$$
(119)
The Disc and the Cone
Up to now, we have mostly solved the problem using the zeta regularization of the Laplacian, and relying on Kronecker’s first limit formula. Only for the case of the torus, we have also performed the calculation of the Robin mass, and verified the prediction of the Morpurgo theorem. In this section we will give the results for a geometry \(\Omega \) in which the calculation of the Robin mass is done with relatively small effort, as the Green function can be guessed through the method of images, while the calculation of the Kronecker mass would require a sum over maxima and minima of Bessel functions.Footnote 10 Let us introduce the notation \(\mathcal D_p(r)\) for the circular sector of radius r and angle \(\frac{2\pi }{p}\), see Fig. 6a,
$$\begin{aligned} \mathcal D_p(r):=\left\{ x\in \mathbb {C}:|x|\le r,\quad 0<\arg x<\frac{2\pi }{p}\right\} . \end{aligned}$$
(120)
The unit area condition implies \(2\pi r^2=p\). We considered the case \(p\in \mathbb {N}\), and we choose periodic boundary conditions in the angular direction: this is equivalent to say that we identify the two radii of the sector, obtaining in this way a cone of height \(r\sqrt{1-p^{-2}}\), see Fig. 6b. This surface is interesting, as it is the first example in our list of a surface with singular curvature, the conical singularity being at the vertex of the cone. We have argued in the introduction that, because of the scaling \(\sim \sqrt{N^{-1}\ln N}\) of the field \(\mu \), we expect that the same theory applies to the flat Euclidean space and to curved manifolds, as long as the curvature is non-singular. The case of surfaces with a finite number of conical singularities would require a different (although feasible) argument, and the verification of our theory on this family of surfaces (as well as the surfaces treated in Sect. 5.5) is an important validation of our predictions.
The Robin’s mass for this case is obtained in Appendix B, and it is equal to
$$\begin{aligned} R_{\mathcal D_p}=-\frac{\ln \pi }{4\pi } +\frac{5p-2}{8\pi } +\frac{\gamma _\text {E}+\psi (1\big /p)}{2\pi }-\frac{\ln p}{4\pi }, \end{aligned}$$
(121)
where \(\phi (z)\) is the digamma function. In particular, for \(\alpha \rightarrow 2\pi \) we recover the case of the unit disc \(\mathcal D\equiv \mathcal D_{1}\):
$$\begin{aligned} R_{\mathcal D}= \frac{1}{\pi } \left( \frac{3}{8} -\frac{\ln \pi }{4} \right) . \end{aligned}$$
(122)
The Kronecker’s mass is readily obtained using Eq. (55).
The Unit Sphere and the Real Projective Sphere
An example of transportation problem on the surface of the sphere \(\mathcal S^2\) has been considered in [42], where the problem of transporting a uniform mass distribution into a set of random points on \(\mathcal S^2\) is analyzed. As an example of applications of our approach to non-flat manifolds, here we consider the problem in our usual setting, i.e., a transportation between two atomic measures of random points. As in the previous cases, the information on the finite-size corrections is related to the spectrum of the Laplace–Beltrami operator on the manifold. It is well known that the eigenfunctions of \(-\Delta \) on the surface of a sphere of radius r are the spherical harmonics \(Y_{l,m}(\theta ,\phi )\) with \(l\in \mathbb {N}\) and \(m\in \mathbb {Z}\) with \(-l\le m\le l\). The corresponding eigenvalues are
$$\begin{aligned} \lambda _{l,m}=\frac{l(l+1)}{r^2}, \end{aligned}$$
(123)
that is, the eigenvalue \({l(l+1)}/{r^2}\) has multiplicity \(2l+1\), for the range of integers \(-l \le m \le l\).
We fix the surface area of the sphere to 1 (taking \(r=(4\pi )^{-1\big /2}\)) and we proceed using the zeta regularization, computing
$$\begin{aligned} Z(s) = \frac{1}{(4 \pi )^s} \sum _{l \ge 1} \frac{2 l+1}{[ l (l+1)]^s}. \end{aligned}$$
(124)
In this case, after some algebra, we are led to use ‘just’ the version of the zeta regularization for the Riemann zeta function (which is much simpler than the Kronecker formula)
$$\begin{aligned} \zeta (s):=\sum _{k\ge 1} \frac{1}{k^s}=\frac{1}{s-1} + \gamma _{\text {E}} + \mathcal O(s-1) \end{aligned}$$
(125)
and we obtain
$$\begin{aligned} Z(s)=\frac{1}{4 \pi (s-1)} - \frac{\ln (4 \pi )}{4\pi } +\frac{\gamma _{\text {E}}}{2 \pi } - \frac{1}{4 \pi } + \mathcal O(s-1). \end{aligned}$$
(126)
The Kronecker mass for the unit sphere is therefore
$$\begin{aligned} K_{\mathcal S^2} = -\frac{1+\ln (4 \pi ) }{4\pi } +\frac{\gamma _{\text {E}}}{2 \pi }. \end{aligned}$$
(127)
Alternatively, we can use one of the regularizations illustrated in Sect. 4. A convenient one is the evaluation of \(W^{(1\big /2)}\), which gives
$$\begin{aligned} \begin{aligned} W^{(1\big /2)}_{\mathcal S^2}(\epsilon )&=r^2\sum _{l \in \mathbb {N}^+}\frac{2l+1}{l(l+1)} \hbox {e}^{-\epsilon r^{-1}\sqrt{l(l+1)}}\\&=r^2 \sum _{l\in \mathbb {N}^+}\left( \frac{1}{l}+\frac{1}{l+1}\right) \hbox {e}^{-\epsilon r^{-1}\sqrt{l(l+1)}}\\&= r^2 \left( 2\ln \frac{r}{\epsilon } - 1\right) +\mathcal {O}(\epsilon ).\end{aligned} \end{aligned}$$
(128)
Recalling that in our case \(r^2=(4\pi )^{-1}\), the final result is
$$\begin{aligned} W^{(1\big /2)}_{\mathcal S^2}(\epsilon ) = -\frac{\ln \epsilon }{2\pi }-\frac{\ln (4\pi )+1}{4\pi }+ \mathcal {O}(\epsilon ), \end{aligned}$$
(129)
that, in light of (59), allows to rederive equation (127).
The spherical lune. The calculation above can be extended to the spherical lune \(\mathcal S_k^2\), a surface on a sphere of radius r, \(4\pi r^2=k\), contained by two half great circles which meet at antipodal points with dihedral angle \(\frac{2\pi }{k}\), see Fig. 7. In Appendix C it is shown that the Kronecker’s mass corresponding to this manifold is
$$\begin{aligned} K_{\mathcal S_k^2}=\frac{k-2-\ln (2\pi k)}{4\pi }+\frac{\gamma _\text {E}}{2\pi }. \end{aligned}$$
(130)
Choosing periodic boundary conditions on the two half great circles, the Kronecker’s mass takes an additional \(-\frac{1}{2\pi }\ln 2\) contribution, and we obtain
$$\begin{aligned} K_{\mathcal S_k^2}=\frac{k-2-\ln (4\pi k)}{4\pi }+\frac{\gamma _\text {E}}{2\pi }, \end{aligned}$$
(131)
that reduces to (127) for \(k=1\), as it should.
The projective sphere. A variation of the problem on the (unit) sphere is the problem on the (unit) real projective sphere \(\mathcal {PS}^2\), that is, the sphere in which antipodal points are identified. The eigenfunctions of the Laplace–Beltrami operator are still the spherical harmonics \(Y_{l,m}(\theta ,\phi )\) with \(l\in \mathbb {N}\) and \(m\in \mathbb {Z}\), \(-l\le m\le l\), but we have to restrict ourselves to eigenfunctions that are invariant under the transformation \((\theta ,\pi ) \mapsto (\pi - \theta ,\phi + \pi )\), i.e., to even values of l. Working on the unit-area manifold accounts to have \(2 \pi r^2=1\). We get
$$\begin{aligned} Z(s) = \frac{1}{(2\pi )^s} \sum _{l \ge 1} \frac{4l+1}{[2l (2l+1)]^s}= \frac{1}{4 \pi (s-1)} - \frac{\ln (2 \pi )}{4\pi }+\frac{\gamma _{\text {E}}}{2 \pi } - \frac{1}{2 \pi } + \mathcal O(s-1) \end{aligned}$$
(132)
so that the Kronecker’s mass is
$$\begin{aligned} K_{\mathcal {PS}^2} = - \frac{\ln (2 \pi )}{4\pi } + \frac{\gamma _E}{2 \pi } - \frac{1}{2 \pi }. \end{aligned}$$
(133)
Using a sharp cut-off instead
$$\begin{aligned} \begin{aligned} W^\text {sharp}_{\mathcal {PS}^2}(\epsilon )&=r^2\sum _{l \in \mathbb {N}^+}\frac{4l+1}{2l(2l+1)}\theta \left( \frac{1}{\epsilon }-\frac{2l(2l+1)}{r^2}\right) \\&=r^2 \sum _{l\in \mathbb {N}^+}\left( \frac{1}{2l}+\frac{1}{2l+1}\right) \theta \left( \frac{1}{2} \sqrt{\frac{r^2}{\epsilon }+\frac{1}{4}}-\frac{1}{4}-l\right) \\&=r^2 \left[ H\left( \frac{r}{\sqrt{\epsilon }}\right) -1\right] +\mathcal {O}(\epsilon )\\&=\frac{1}{2 \pi }\left( -\frac{1}{2}\ln \epsilon -\frac{\ln (2 \pi )}{2}+\gamma _\text {E}-1 +\mathcal {O}(\epsilon )\right) \end{aligned} \end{aligned}$$
(134)
which, again by using (63) and (65), allows to rederive equation (133).