This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Vortex pairs and dipoles on closed surfaces

Björn Gustafsson1
Abstract

We set up general equations of motion for point vortex systems on closed Riemannian surfaces, allowing for the case that the sum of vorticities is not zero and there hence must be counter-vorticity present. The dynamics of global circulations which is coupled to the dynamics of the vortices is carefully taken into account.

Much emphasis is put to the study of vortex pairs, having the Kimura conjecture in focus. This says that vortex pairs move, in the dipole limit, along geodesic curves, and proofs for it have previously been given by S. Boatto and J. Koiller by using Gaussian geodesic coordinates. In the present paper we reach the same conclusion by following a slightly different route, leading directly to the geodesic equation with a reparametrized time variable.

In a final section we explain how vortex motion in planar domains can be seen as a special case of vortex motion on closed surfaces, and in two appendices we give some necessary background on affine and projective connections.

Keywords: Point vortex, vortex pair, vortex dipole, geodesic curve, affine connection, projective connection, Robin function, Hamiltonian function, symplectic form, Green function.

MSC: 76B47, 53B05, 53B10, 30F30

Acknowledgements: The author has been inspired by discussions with Stefanella Boatto, Jair Koiller, Nikolai Nadirashvili, and Jesper Göransson.

11footnotetext: Department of Mathematics, KTH, 100 44, Stockholm, Sweden.
Email: [email protected]

1 Introduction

This paper extends previous results in [24] on multiple point vortex motion on closed Riemannian surfaces of arbitrary genus to cases in which the sum of the vorticities is not zero, and therefore a counter-vorticity must be present. This occurs for example when there is only one point vortex. For reasons dictated by a Hodge decomposition, the counter-vorticity is naturally chosen to be uniformly distributed on the surface.

Special for our investigations is that we, in the case of higher genus, carefully take the circulation around the holes into account and make sure that the Kelvin law of preservation of circulations is satisfied. This was done already in [24], but in the case of a counter-vorticity being present one has to take some extra steps. It is natural to choose fixed closed curves on the surface to make up a basis for the first homology group, but the circulation around such curves will in general not be conserved in time. In the present paper we handle the so arising problem by considering these circulations as free variables in the phase space, in addition to the locations of the vortices. If there are nn vortices and the genus of the surface is g the phase space will thus have real dimension 2n+2g2n+2\texttt{g}.

The Hamiltonian function will, as usual, be a renormalized kinetic energy for the flow, see (4.26), but in the presence of a counter-vorticity the symplectic form has to be accordingly adapted, see (5.1). With this done the dynamical equations come out to be the expected ones (Theorem 5.1). After having set up the vortex dynamics in general (Section 4 and 5) and after a short discussion of the single vortex case in genus zero (Section 6), we turn in Section 7 to the question of motion of vortex pairs, i.e. systems consisting of two point vortices of equal strength but rotating in opposite direction. In this case there is no counter-vorticity. The conjecture of Y. Kimura [28], saying that a vortex pair in the dipole limit moves along a geodesic curve, has been a major source of inspiration for the present paper. The same applies for papers [31, 3] by S. Boatto, J. Koiller. These latter papers actually contain a proof of Kimura’s conjecture, and in the present paper we try to clarify the situation further by connecting the dipole trajectory directly to the geodesic equation associated to the metric.

Throughout this paper we think of a Riemannian surface (i.e. a Riemannian manifold of dimension two) as a Riemann surface provided with a metric which is compatible with the conformal structure. Therefore we are lead into more or less classical function theory on Riemann surfaces. Our treatment is based entirely on local holomorphic coordinates. For example, the logarithmic singularities in the stream function will look like log|zw|\log|z-w| near a vortex, zz and ww being complex coordinates, ww representing the location of a vortex. When such a singular term is removed from a local series expansion (with respect to zz) around ww in the necessary renormalization process and zz is set equal to ww, then the quantity which remains behaves no longer as a function under changes of coordinates. Instead it is a kind of “connection”. For such reasons we run into differential geometric considerations involving different kinds of connections: affine connections (or 11-connections) and projective, or Schwarzian, connections (22-connections). The metric itself can be viewed as the exponential of a 0-connection. The necessary definitions and background material are summarized in two appendices, Sections 10 and 11. Affine connections appear in many areas of mathematical physics, and as a mathematical tool they show up in expressions for covariant derivatives, and in particular in the equation (10.5) for geodesic curves. Also projective connections are regularly used, for example in conformal field theory.

The Kimura conjecture and other matters related to dipole motion has been discussed also in [37, 38, 41, 30, 32], and from slightly different points of view in [7, 33, 27, 6]. As mentioned, the analysis in the present paper leads to a new confirmation of Kimura’s conjecture, although in a rather weak form: expanding all quantities in Taylor series and keeping only the leading terms, the dynamical equations for a vortex pair reduce to the geodesic equation in the dipole limit. See Theorem 7.1. It seems likely that stronger forms of the conjecture (convergence of the trajectories, for example) can be formulated and proved, but we leave such matters for possible future investigations.

It may be remarked that Kimura’s conjecture is counter-intuitive. The reason to think so is that vortex motion in general is governed by global laws on the manifold, like the structure of Green functions and other harmonic functions, whereas the geometry of geodesics is an entirely local matter. If the geometry of the manifold is changed at one place then this will not affect what geodesics look like at another place. It will however change the structure of Green functions and the general vortex dynamics everywhere. The solution of this paradox is that vortex dipoles are highly singular. A vortex pair is partly governed by the global harmonic structure, but in the limit, when the vortex pair becomes a dipole, this harmonic structure is completely overruled by the differential geometric structure. In that limit, all terms in the dynamical equations containing harmonic functions become negligible compared to those terms depending on the metric only. See Section 7.

In Section 8 we make some attempts to treat vortex dipoles directly by starting from the dynamical equation for a single vortex and taking the distributional derivative with respect to the coordinate ww which gives the location of the vortex. The local stream function for the flow near a single vortex at ww is something like (we ignore constant factors below)

ψ(z)=log|zw|+regular terms,\psi(z)=\log|z-w|+\text{regular terms},

and the corresponding flow, when considered as a one-form (or covariant vector field) looks like

ν(z)=dzzw+complex conjugate+regular terms.\nu(z)=\frac{dz}{z-w}+\text{complex conjugate}+\text{regular terms}.

The term named “complex conjugate” is inserted to make possible for ν\nu to be real-valued. For the dipole we accordingly have

dwν(z)=dzdw(zw)2+complex conjugates+regular terms.d_{w}\nu(z)=\frac{dzdw}{(z-w)^{2}}+\text{complex conjugates}+\text{regular terms}.

The differential dwdw shall then be thought of as containing information of the orientation of the dipole.

The constant term (i.e. the first term in the Taylor series of the regular terms) are in the above three expressions a 0-connection, a 11-connection, and a 22-connection, respectively (again up to constant factors). See Lemma 11.1 for a precise statement. The 0-connection for the stream function is the Routh’s stream function, or Robin function when ψ\psi is thought of as a Green function. And the 11-connection for ν\nu is what gives the speed of the vortex, after subtraction (in case of a curved manifold) of the affine connection coming from the metric. The result will then be a covariant vector field.

For dwν(z)d_{w}\nu(z) the constant term is similarly a 22-connection, or projective connection, but it is more unclear what influence it has on the motion of the dipole. Indeed, the dipole singularity is very strong and it has a definite direction at ww, and it seems that all finite terms should be negligible in comparison. Therefore it is difficult to figure out any dynamics from this picture, only that the dipole should move with infinite speed in the direction dictated by the singularity, presumably (and as have been confirmed) along a geodesic with respect to the metric.

In the last section (not counting the appendices) of the paper we discuss briefly how point vortex motion in planar domains can be considered as a special case of vortex motion on surfaces. This is done by doubling the domain to a compact Riemann surface (the Schottky double), which is a standard tool as far as the conformal structure is concerned. What is special in our case is that we have to take the metric structure into account, and this becomes non-smooth in the doubling procedure. An interesting observation is that the boundary of the domain becomes a geodesic curve with respect to the natural planar metric (on each of the halves) of the Schottky double.

In general the present paper is, in addition to the papers by S. Boatto and J. Koiller already mentioned, much in spirit of papers [37, 38, 18, 19, 49] by S. G. Llewellyn Smith, R. J. Nagem, C. Grotta Ragazzo, H. Viglioni, and Q. Wang. The paper [4] by A. Bogatskiy contains ideas concerning the higher genus case which are somewhat similar to those in the present paper, see in particular Section A.3 in [4]. We wish to mention also the work [5] by A. V. Borisov and I. S. Mamaev, and the very early article [16] (cited in [5, 38]) by A. A. Fridman and P. Ya. Polubarinova. That work, from 1928, may be one of the first papers on motion of dipoles and other higher singularities. Fridman was a famous cosmologist and Pelageya Yakolevna Polubarinova (or Polubarinova-Kochina, after marriage) was a young (at that time) disciple of Fridman. This remarkable woman wrote her first paper in 1924, and was scientifically active throughout her life. Her published papers and books span a period of 75 years, the last paper appearing the same year as she died, one hundred years of age. See [48] for a short biography related to her work on Hele-Shaw flows.

2 Fluid dynamics on a Riemannian surface

2.1 General notations and assumptions

We consider the dynamics of a non-viscous incompressible fluid on a compact Riemannian manifold of dimension two. We follow the treatments in [46, 1, 15] in the respects of treating the fluid velocity field as a one-form and in extensively using the Lie derivative. These sources also provide the standard notions and notations for differential geometry to be used. Other treatises in fluid dynamics, suitable for our purposes, are [39, 40].

Traditionally, non-viscous fluid dynamics is discussed in terms of the fluid velocity vector field 𝐯{\bf v}, the density ρ\rho, and the pressure pp of the fluid. The basic equations are the equation of continuity (expressing conservation of mass) and Euler’s equation (conservation of momentum):

ρt+(ρ𝐯)=0,\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho{\bf v})=0, (2.1)
𝐯t+(𝐯)𝐯=1ρp.\frac{\partial{\bf v}}{\partial t}+({\bf{v}}\cdot\nabla){\bf v}=-\frac{1}{\rho}\nabla p. (2.2)

These are to be combined with a constitutive law giving a relationship between pp and ρ\rho.

The above equations are already simplified versions of the general Navier-Stokes equations (which allow for viscosity terms), but we shall simplify further by working only in two dimensions and by taking the constitutive law to be the simplest possible:

ρ=1.\rho=1. (2.3)

Thus ρ\rho disappears from discussion, and the equation of continuity becomes

𝐯=0.\nabla\cdot{\bf v}=0. (2.4)

One can get rid also of pp, because when ρ\rho is constant then pp appears only as a scalar potential in Euler’s equation, this equation effectively saying that

𝐯t+(𝐯)𝐯=(something),\frac{\partial{\bf v}}{\partial t}+({\bf{v}}\cdot\nabla){\bf v}=\nabla({\rm something}), (2.5)

where this ``something"=p{\rm``something"}=p  afterwards can be recovered, up to a (time dependent) constant.

So everything is extremely simple (in theory), and even more so in two dimensions. As is well-known [44], every (oriented) Riemannian manifold of dimensions two can be made into a Riemann surface by choosing isothermal local coordinates (x,y)=(x1,x2)(x,y)=(x^{1},x^{2}), by which z=x+iyz=x+\mathrm{i}y becomes a holomorphic coordinate and the metric takes the form

ds2=λ(z)2|dz|2=λ(z)2(dx2+dy2)ds^{2}=\lambda(z)^{2}|dz|^{2}=\lambda(z)^{2}(dx^{2}+dy^{2}) (2.6)

with λ>0\lambda>0. Thus the metric tensor gijg_{ij}, as appearing in general in ds2=gijdxidxjds^{2}=g_{ij}dx^{i}dx^{j}, becomes gij=λ2δijg_{ij}=\lambda^{2}\delta_{ij} in these coordinates. It turns out to be convenient to work with the fluid velocity 11-form (or covariant vector)

ν=νxdx+νydy\nu=\nu_{x}\,dx+\nu_{y}\,dy

rather than with the corresponding vector field, which then becomes

𝐯=1λ2(νxx+νyy).{\bf v}=\frac{1}{\lambda^{2}}(\nu_{x}\,\frac{\partial}{\partial x}+\nu_{y}\,\frac{\partial}{\partial y}).

The Hodge star operator takes 0-forms into 22-forms and vice versa, and takes 11-forms to 11-forms. It acts on basic differential forms as

1=λ2(dxdy)=vol=the volume (area) form,*1=\lambda^{2}(dx\wedge dy)={\rm vol}=\text{the volume (area) form},
dx=dy,dy=dx,(dxdy)=λ2.*dx=dy,\quad*dy=-dx,\quad*(dx\wedge dy)=\lambda^{-2}.

Thus, for 11-forms,

ν=νydx+νxdy,*\nu=-\nu_{y}\,dx+\nu_{x}\,dy,

which can be interpreted as a rotation by ninety degrees of the corresponding vector.

In addition to the Hodge star it is useful to introduce the Lie derivative 𝐯\mathcal{L}_{\bf v} of a vector field 𝐯{\bf v}. When acting on differential forms this is related to interior derivation (or “contraction”) i(𝐯)i({\bf v}) and exterior derivation dd by the homotopy formula

𝐯=di(𝐯)+i(𝐯)d.\mathcal{L}_{\bf{v}}=d\circ i({\bf{v}})+i({\bf{v}})\circ d. (2.7)

The Hodge star and interior derivation are related by

i(𝐯)vol=ν,i({\bf v}){\rm vol}=*\nu, (2.8)

where 𝐯{\bf v} and ν\nu are linked via the metric tensor as above. Thus

dν=d(i(𝐯)vol)=𝐯(vol)=(𝐯)vol.d*\nu=d(i({\bf v}){\rm vol})=\mathcal{L}_{\bf v}({\rm vol})=(\nabla\cdot{\bf v}){\rm vol}.

See [15] for this identity, and for further details in general. We conclude that (2.4) is equivalent to the statement that ν*\nu is a closed form:

dν=0.d*\nu=0. (2.9)

Locally we can therefore write

ν=dψ*\nu=d\psi (2.10)

for some (locally defined) stream function ψ\psi. The vorticity 22-form for a fluid is, in terms of the flow 11-form ν,\nu,

ω=dν=(νyxνxy)dxdy.\omega=d\nu=(\frac{\partial\nu_{y}}{\partial x}-\frac{\partial\nu_{x}}{\partial y})\,dx\wedge dy. (2.11)

The Euler equation (2.2) can be written

(t+𝐯)(ν)=d(12|𝐯|2p).(\frac{\partial}{\partial t}+\mathcal{L}_{\bf v})({\bf\nu})=d(\frac{1}{2}|{\bf v}|^{2}-{p}). (2.12)

Note that the left member involves the fluid velocity both as a vector field and as a form. The combination t+𝐯\frac{\partial}{\partial t}+\mathcal{L}_{\bf v} can be viewed as a counterpart, for forms, to the more traditional “convective derivative” (or material derivative) t+𝐯\frac{\partial}{\partial t}+{\bf{v}}\cdot\nabla, implicitly used in (2.5), for vector fields. However, the two derivatives are not the same. The equivalence between (2.12) and (2.2) (when ρ=1\rho=1) is a consequence of the identity

𝐯ν=d(12|𝐯|2)+(𝐯)ν,\mathcal{L}_{\bf v}\nu=d(\frac{1}{2}|{\bf v}|^{2})+({\bf{v}}\cdot\nabla){\nu},

which can be directly verified by coordinate computations. See [46, 15] for details.

Since the pressure pp does not appear in any other equation, the Euler equation on the form (2.12) only expresses, in analogy with (2.5), that the left member is an exact 11-form:

(t+𝐯)(ν)=exact=dϕ.(\frac{\partial}{\partial t}+\mathcal{L}_{\bf v})({\bf\nu})={\rm exact}=d\phi. (2.13)

From the scalar ϕ\phi the pressure pp then can be recovered via

ϕ=12|𝐯|2p+constant.\phi=\frac{1}{2}|{\bf v}|^{2}-{p}+{\rm constant}. (2.14)

On acting by dd on (2.13), the local form of Helmholtz-Kirchhoff-Kelvin law of conservation of vorticity follows:

(t+𝐯)(ω)=0.(\frac{\partial}{\partial t}+\mathcal{L}_{\bf{v}})(\omega)=0. (2.15)

Since the right member of (2.13) is not only a closed differential form, but even an exact form, a stronger, global, version of the Helmholtz-Kirchhoff-Kelvin law actually follows. This can be expressed by saying that

ddtγ(t)ν=0\frac{d}{dt}\oint_{\gamma(t)}\nu=0

for any closed curve γ(t)\gamma(t) which moves with the fluid.

3 Green functions and harmonic forms

3.1 The one point Green function

In the sequel MM will be a closed (compact) Riemann surface provided with a Riemannian metric on the form (2.6).

Given a 22-form ω\omega on MM one can immediately obtain a corresponding Green potential GωG^{\omega} by Hodge decomposition, i.e. by orthogonal decomposition in the Hilbert space of square integrable 22-forms. The inner product for such forms is

(ω1,ω2)2=Mω1ω2.(\omega_{1},\omega_{2})_{2}=\int_{M}\omega_{1}\wedge*\omega_{2}. (3.1)

The given 22-form ω\omega then decomposes as the orthogonal decomposition of an exact form and a harmonic form:

ω=d(something)+harmonic.\omega=d(\rm something)+{\rm harmonic}.

This Hodge decomposition can more precisely be written as

ω=ddGω+constantvol,\omega=-d*dG^{\omega}+{\rm constant}\cdot{\rm vol}, (3.2)

where GωG^{\omega} is normalized to be orthogonal to all harmonic 22-forms, namely satisfying

MGωvol=0.\int_{M}G^{\omega}\,{\rm vol}=0. (3.3)

The constant in (3.2) necessarily equals the mean value of ω\omega,

constant=1VMω,{\rm constant}=\frac{1}{V}\int_{M}\omega, (3.4)

where VV denotes the total volume (=area) of MM:

V=Mvol.V=\int_{M}{\rm vol}.

When ω\omega is exact, as in (2.11), the second term in the right member of (3.2) disappears, hence

ddGω=ω-d*dG^{\omega}=\omega (3.5)

in this case. In the other extreme, when ω\omega is harmonic, i.e. is a constant multiple of vol{\rm vol}, the first term disappears. Indeed, the whole Green function disappears:

Gvol=0.G^{{\rm vol}}=0. (3.6)

For 11-forms the inner product has the same expression as for 22-forms:

(ν1,ν2)1=Mν1ν2.(\nu_{1},\nu_{2})_{1}=\int_{M}\nu_{1}\wedge*\nu_{2}. (3.7)

If ν\nu represents a fluid velocity, then (ν,ν)1(\nu,\nu)_{1} has the interpretation of being (twice) the kinetic energy of the flow. For a function (potential) uu we consider the Dirichlet integral (du,du)1(du,du)_{1} to be its energy. Thus constant functions have no energy.

The energy (ω,ω)\mathcal{E}(\omega,\omega) of any 22-form ω\omega is defined to be the energy of its Green potential GωG^{\omega}. Thus, for the mutual energy,

(ω1,ω2)=(dGω1,dGω2)1=MGω1ω2.\mathcal{E}(\omega_{1},\omega_{2})=(dG^{\omega_{1}},dG^{\omega_{2}})_{1}=\int_{M}G^{\omega_{1}}\wedge\omega_{2}. (3.8)

It follows, from (3.3) for example, that the volume form has no energy:

(vol,vol)=0.\mathcal{E}({\rm vol},{\rm vol})=0.

It was tacitly assumed above that the forms under discussion belong to the L2L^{2}-space defined by the inner product. However, the mutual energy sometimes extends to circumstances in which source distributions have infinite energy. This applies in particular to the Dirac current δa\delta_{a}, which we consider as a 22-form with distributional coefficient, namely defined by the property that

Mδaφ=φ(a)\int_{M}\delta_{a}\wedge\varphi=\varphi(a)

for every smooth function φ\varphi. Certainly δa\delta_{a} has infinite energy, but if aba\neq b, then (δa,δb)\mathcal{E}(\delta_{a},\delta_{b}) is still finite and has a natural interpretation: it is the (one-point) Green function, or “monopole” Green function:

G(a,b)=Gδa(b)=(δa,δb).G(a,b)=G^{\delta_{a}}(b)=\mathcal{E}(\delta_{a},\delta_{b}). (3.9)

Here the first equality can be taken as a definition, and then the second equality follows on using (3.2), (3.3):

(δa,δb)\displaystyle\mathcal{E}(\delta_{a},\delta_{b}) =MdGδadGδb=MGδaddGδb\displaystyle=\int_{M}dG^{\delta_{a}}\wedge*dG^{\delta_{b}}=-\int_{M}G^{\delta_{a}}\wedge d*dG^{\delta_{b}}
=MGδa(δb1Vvol)=Gδa(b)=G(a,b).\displaystyle=\int_{M}G^{\delta_{a}}\wedge\left(\delta_{b}-\frac{1}{V}\,{\rm vol}\right)=G^{\delta_{a}}(b)=G(a,b).

This shows in addition that G(a,b)G(a,b) is symmetric.

Changing now notations from a,ba,b to z,wz,w, where later ww will have the role of being the location of a point vortex, the Green function G(z,w)G(z,w), as a function of zz, has just one pole (at z=wz=w). Its Laplacian, as a 22-form, is then

ddG(,w)=δw1Vvol.-d*dG(\cdot,w)=\delta_{w}-\frac{1}{V}{\rm{vol}}. (3.10)

It is interesting to notice that among the two terms in the right member of (3.10), one has infinite energy and one has zero energy ((δw,δw)=+\mathcal{E}(\delta_{w},\delta_{w})=+\infty, (vol,vol)=0\mathcal{E}({\rm vol},{\rm vol})=0).

Remark 3.1.

It is more common to treat the Dirac delta and the Laplacian (denoted Δ\Delta) as “densities” with respect to the volume form. However, we find our usage convenient. In any case, the relationships are

δa=(delta′′function′′)vol,ddφ=(Δφ)vol.\delta_{a}=({\rm delta\,\,^{\prime\prime}function^{\prime\prime}})\,{\rm vol},\quad d*d\varphi=(\Delta\varphi)\,{\rm vol}.
Remark 3.2 (Notational remark).

Letters like zz, ww will have the double roles of being complex-valued local coordinates on parts of the Riemann surface and of denoting the points on the surface for which the coordinates have the values in question. A more formal treatment could use, for example, PMP\in M for a point and z(P)z(P)\in{\mathbb{C}} for the corresponding coordinate value.

Remark 3.3 (Real versus complex notation).

For future needs we wish to clarify the relationship between real and complex coordinates in the context of tangent and cotangent vectors.

Let z=x+iyz=x+\mathrm{i}y be a complex coordinate on MM and consider a curve tz(t)t\mapsto z(t) in MM (for example the trajectory of a vortex). Set z˙=dz/dt\dot{z}={dz}/{dt}, and similarly for x˙\dot{x} and y˙\dot{y}, so that z˙=x˙+iy˙\dot{z}=\dot{x}+\mathrm{i}\dot{y}. The velocity of this moving point is primarily to be considered as a vector in the (real) tangent space of MM (at the point under consideration). This gives, in the picture of viewing tangent vectors as derivations, the velocity vector

𝐕=x˙x+y˙y=z˙z+z¯˙z¯,{\bf V}=\dot{x}\frac{\partial}{\partial x}+\dot{y}\frac{\partial}{\partial y}=\dot{z}\frac{\partial}{\partial z}+\dot{\bar{z}}\frac{\partial}{\partial\bar{z}}, (3.11)

where it is understood that z¯˙=z˙¯\dot{\bar{z}}=\overline{\dot{z}}.

The real tangent space used above can in a next step be complexified, which means that one breaks the connection between z˙\dot{z} and z¯˙\dot{\bar{z}} and consider them as independent complex variables. Equivalently, one allows x˙\dot{x} and y˙\dot{y} to be complex-valued. The so obtained complex tangent space can be decomposed as a direct sum of its holomorphic and anti-holomorphic subspaces, and then

proj:𝐕=z˙z+z¯˙z¯z˙z{\rm proj}:\quad{\bf V}=\dot{z}\frac{\partial}{\partial z}+\dot{\bar{z}}\frac{\partial}{\partial\bar{z}}\mapsto\dot{z}\frac{\partial}{\partial z}

is a natural and useful identification of the real tangent space with the holomorphic part of the complex tangent space. See [17], in particular Section 2 of Chapter 0, for further discussions. With this identification z˙\dot{z} represents the velocity of z(t)z(t). Still one need to keep in mind that z˙\dot{z} is just a complex number and that it is rather the preimage under proj{\rm proj} above that is the true velocity, as a real tangent vector.

The vector 𝐕{\bf V} above correspond, via the metric, to the covector

λ2(x˙dx+y˙dy)=λ22(z¯˙dz+z˙dz¯),{\lambda^{2}}(\dot{x}dx+\dot{y}dy)=\frac{\lambda^{2}}{2}(\dot{\bar{z}}dz+\dot{{z}}d\bar{z}), (3.12)

where λ=λ(x,y)=λ(z)\lambda=\lambda(x,y)=\lambda(z) (depending on context). Note that

12z¯˙=12(x˙iy˙),12z˙=12(x˙+iy˙),\frac{1}{2}\dot{\bar{z}}=\frac{1}{2}(\dot{x}-\mathrm{i}\dot{y}),\quad\frac{1}{2}\dot{{z}}=\frac{1}{2}(\dot{x}+\mathrm{i}\dot{y}),

as coefficients of dzdz and dz¯d\bar{z} (respectively) have similar roles (and signs) as the Wirtinger derivatives /z\partial/\partial z, /z¯\partial/\partial\bar{z}. For a covector ν\nu in general we therefore define

νz=12(νxiνy),νz¯=12(νx+iνy),\nu_{z}=\frac{1}{2}(\nu_{x}-\mathrm{i}\nu_{y}),\quad\nu_{\bar{z}}=\frac{1}{2}(\nu_{x}+\mathrm{i}\nu_{y}),

so that

ν=νxdx+νydy=νzdz+νz¯dz¯.\nu=\nu_{x}dx+\nu_{y}dy=\nu_{z}dz+\nu_{\bar{z}}d\bar{z}.

The corresponding (contravariant) vector is then, as in our fluid dynamical contexts,

𝐯=1λ2(νxx+νyy)=1λ2(νzz+νz¯z¯).{\bf v}=\frac{1}{\lambda^{2}}\big{(}\nu_{x}\frac{\partial}{\partial x}+\nu_{y}\frac{\partial}{\partial y}\big{)}=\frac{1}{\lambda^{2}}\big{(}\nu_{z}\frac{\partial}{\partial z}+\nu_{\bar{z}}\frac{\partial}{\partial\bar{z}}\big{)}.

3.2 Harmonic one-forms and period relations

For later use we record the formulas (differentiation with respect to zz)

dG(z,w)=G(z,w)zdz+G(z,w)z¯dz¯=2Re(G(z,w)zdz),dG(z,w)=\frac{\partial G(z,w)}{\partial z}dz+\frac{\partial G(z,w)}{\partial\bar{z}}d\bar{z}=2\operatorname{Re}\big{(}\frac{\partial G(z,w)}{\partial z}dz\big{)},\qquad
dG(z,w)=iG(z,w)zdz+iG(z,w)z¯dz¯=2Im(G(z,w)zdz).*dG(z,w)=-\mathrm{i}\frac{\partial G(z,w)}{\partial z}dz+\mathrm{i}\frac{\partial G(z,w)}{\partial\bar{z}}d\bar{z}=2\operatorname{Im}\big{(}\frac{\partial G(z,w)}{\partial z}dz\big{)}. (3.13)

If γ\gamma is any closed oriented curve in MM then clearly γ𝑑G(,w)=0\oint_{\gamma}dG(\cdot,w)=0, while the conjugate period defines a function

Uγ(w)=γdG(,w)=γ(iG(z,w)zdz+iG(z,w)z¯dz¯)U_{\gamma}(w)=\oint_{\gamma}*dG(\cdot,w)=\oint_{\gamma}\big{(}-\mathrm{i}\frac{\partial G(z,w)}{\partial z}dz+\mathrm{i}\frac{\partial G(z,w)}{\partial\bar{z}}d\bar{z}\big{)} (3.14)
=2iγG(z,w)z𝑑z=2iγG(z,w)z¯𝑑z¯,=-2\mathrm{i}\oint_{\gamma}\frac{\partial G(z,w)}{\partial z}dz=2\mathrm{i}\oint_{\gamma}\frac{\partial G(z,w)}{\partial\bar{z}}d\bar{z}, (3.15)

which, away from γ\gamma, is harmonic in ww and makes a unit additive jump as ww crosses γ\gamma from the left to the right. The harmonicity of Uγ(wU_{\gamma}(w) is perhaps not obvious from outset since G(,w)G(\cdot,w) is not itself harmonic, but the deviation from harmonicity, namely the extra term in (3.10), is independent of ww and therefore disappears under differentiation.

Differentiating (3.14) with respect to ww gives

dUγ(w)=2iγ2G(z,w)zw𝑑z𝑑w2iγ2G(z,w)zw¯𝑑z𝑑w¯,dU_{\gamma}(w)=-2\mathrm{i}\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial w}d{z}\otimes dw-2\mathrm{i}\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial\bar{w}}d{z}\otimes d\bar{w},
dUγ(w)=2γ2G(z,w)zwdzdw+2γ2G(z,w)zw¯dzdw¯*dU_{\gamma}(w)=-2\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial w}d{z}\otimes dw+2\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial\bar{w}}d{z}\otimes d\bar{w}\quad

(integration with respect ot zz, Hodge star and dd with respect to ww). We have written out a tensor product sign to emphasize that the product between the dzdz and dwdw is not a wedge product. Adding and subtracting we obtain the analytic (respectively anti-analytic) differentials

dUγ(w)+idUγ(w)=2Uγ(w)wdw=4iγ2G(z,w)zw𝑑z𝑑w,dU_{\gamma}(w)+\mathrm{i}*dU_{\gamma}(w)=2\frac{\partial U_{\gamma}(w)}{\partial w}dw=-4\mathrm{i}\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial w}d{z}\otimes dw, (3.16)
dUγ(w)idUγ(w)=2Uγ(w)w¯dw¯=4iγ2G(z,w)zw¯𝑑z𝑑w¯.dU_{\gamma}(w)-\mathrm{i}*dU_{\gamma}(w)=2\frac{\partial U_{\gamma}(w)}{\partial\bar{w}}d\bar{w}=-4\mathrm{i}\oint_{\gamma}\frac{\partial^{2}G(z,w)}{\partial{z}\partial\bar{w}}d{z}\otimes d\bar{w}. (3.17)

Let now {α1,,αg,β1,,βg}\{\alpha_{1},\dots,\alpha_{\texttt{g}},\beta_{1},\dots,\beta_{\texttt{g}}\} be representing cycles for a canonical homology basis for MM such that each βj\beta_{j} intersects αj\alpha_{j} once from the right to the left and no other intersections occur (see [12] for details). Then there are corresponding harmonic differentials dUαjdU_{\alpha_{j}}, dUβjdU_{\beta_{j}} obtained on choosing γ=αj,βj\gamma=\alpha_{j},\beta_{j} in the above construction, and these constitute a basis of harmonic differentials associated with the chosen homology basis. Precisely we have, for k,j=1,,gk,j=1,\dots,\texttt{g},

αk(dUβj)=δkj,βk(dUβj)=0,\oint_{\alpha_{k}}(-dU_{\beta_{j}})=\delta_{kj},\quad\oint_{\beta_{k}}(-dU_{\beta_{j}})=0, (3.18)
αk𝑑Uαj=0,βk𝑑Uαj=δkj.\oint_{\alpha_{k}}dU_{\alpha_{j}}=0,\qquad\oint_{\beta_{k}}dU_{\alpha_{j}}=\delta_{kj}. (3.19)

The Kronecker deltas, δkj\delta_{kj}, come from the discontinuity (jump) properties of UγU_{\gamma} mentioned after (3.14).

The integrals of the basic harmonic differentials along non-closed curves become periods of two point Green functions:

ba𝑑Uαj=αjdGδaδb,\int_{b}^{a}dU_{\alpha_{j}}=\oint_{\alpha_{j}}*dG^{\delta_{a}-\delta_{b}}, (3.20)
ba𝑑Uβj=βjdGδaδb.\int_{b}^{a}dU_{\beta_{j}}=\oint_{\beta_{j}}*dG^{\delta_{a}-\delta_{b}}. (3.21)

Here the integration in the left member is to be performed along a path that does not intersect the curve in the right member. These formulas are immediate from the definition (3.14) of UγU_{\gamma}.

4 Energy renormalization and the Hamiltonian

4.1 The renormalized kinetic energy

When ν\nu is the flow 11-form of an incompressible fluid, the equation of continuity says that dν=0d*\nu=0 (see (2.9)). The vorticity 22-form is ω=dν\omega=d\nu, and (3.5) holds. Thus d(ν+dGω)=0d(\nu+*dG^{\omega})=0, and setting

η=ν+dGω\eta=\nu+*dG^{\omega} (4.1)

it follows that η\eta is harmonic: dη=0=dηd\eta=0=d*\eta. Locally we can write η=dψ0*\eta=d\psi_{0} for some harmonic function ψ0\psi_{0}. Then

ψ=Gω+ψ0\psi=G^{\omega}+\psi_{0} (4.2)

becomes a locally defined stream function for the flow, so that

ν=dψ=dGω+dψ0.*\nu=d\psi=dG^{\omega}+d\psi_{0}.

Different local choices of ψ\psi may differ by additive time dependent constants. The roles of the additional terms η\eta and ψ0\psi_{0}, complementing the contribution from the Green function, will be somewhat clarified in the context of planar vortex motion and the hydrodynamic Green function in Section 9.

The relation (4.1) written on the form ν=ηdGω\nu=\eta-*dG^{\omega} is an orthogonal decomposition of the flow 11-form ν\nu with respect to the inner product (3.7). Indeed,

(η,dGω)1=MηdGω=MdηGω=0,(\eta,-*dG^{\omega})_{1}=\int_{M}\eta\wedge dG^{\omega}=-\int_{M}d\eta\wedge G^{\omega}=0,

since η\eta is harmonic. It follows that (twice) the total (kinetic) energy (ν,ν)1(\nu,\nu)_{1} of the flow is given by

2(ν,ν)1=Mνν=i2M(ν+iν)(νiν)=2(\nu,\nu)_{1}=\int_{M}\nu\wedge*\nu=\frac{\mathrm{i}}{2}\int_{M}(\nu+\mathrm{i}*\nu)\wedge(\nu-\mathrm{i}*\nu)=
=MdGωdGω+Mηη=(ω,ω)+Mηη==\int_{M}dG^{\omega}\wedge*dG^{\omega}+\int_{M}\eta\wedge*\eta=\mathcal{E}(\omega,\omega)+\int_{M}\eta\wedge*\eta=
=MGωω+j=1g(αjηβjηαjηβjη).=\int_{M}G^{\omega}\wedge\omega+\sum_{j=1}^{\texttt{g}}(\oint_{\alpha_{j}}\eta\oint_{\beta_{j}}*\eta-\oint_{\alpha_{j}}*\eta\oint_{\beta_{j}}\eta).

We shall now specialize on the point vortex case, with vortices of strengths Γk\Gamma_{k} located at points wkMw_{k}\in M (k=1,,nk=1,\dots,n). However, we shall not assume that these strengths add up to zero. Hence the sum

Γ=k=1nΓk\Gamma=\sum_{k=1}^{n}\Gamma_{k} (4.3)

may be non-zero and there will then be a compensating uniform counter vorticity. The total vorticity ω=dν\omega=d\nu, which satisfies Mω=0\int_{M}\omega=0, appears as the right member in

ddGω=ω=k=1nΓkδwkΓVvol.-d*dG^{\omega}=\omega=\sum_{k=1}^{n}\Gamma_{k}\delta_{w_{k}}-\frac{\Gamma}{V}{\rm vol}. (4.4)

Explicitly we have, on recalling (3.6),

Gω(z)=k=1nΓkG(z,wk).G^{\omega}(z)=\sum_{k=1}^{n}\Gamma_{k}{G(z,w_{k})}. (4.5)

On using (3.13) the conjugate α\alpha-periods of GωG^{\omega} can be expressed as

αjdGω=k=1n2iΓkαjG(z,wk)z𝑑z.\oint_{\alpha_{j}}*d{G^{\omega}}=-\sum_{k=1}^{n}2\mathrm{i}\Gamma_{k}\oint_{\alpha_{j}}\frac{\partial G(z,w_{k})}{\partial z}dz. (4.6)

By differentiation and using also (3.15), (3.16) we have

wkαjdGω=ΓkUαj(wk)wk=iΓkUαj(wk)wk,\frac{\partial}{\partial w_{k}}\oint_{\alpha_{j}}*d{G^{\omega}}=\Gamma_{k}\frac{\partial U_{\alpha_{j}}(w_{k})}{\partial w_{k}}=\mathrm{i}\Gamma_{k}\frac{\partial U^{*}_{\alpha_{j}}(w_{k})}{\partial w_{k}}, (4.7)
wk(αjdGω)dwk=Γk2(dUαj(wk)+idUαj(wk)).\frac{\partial}{\partial w_{k}}\big{(}\oint_{\alpha_{j}}*d{G^{\omega}}\big{)}dw_{k}=\frac{\Gamma_{k}}{2}(dU_{\alpha_{j}}(w_{k})+\mathrm{i}*dU_{\alpha_{j}}(w_{k})). (4.8)

Here UαjU^{*}_{\alpha_{j}} denotes a harmonic conjugate of UαjU_{\alpha_{j}}, whereby dUαj=d(Uαj)*dU_{\alpha_{j}}=d(U^{*}_{\alpha_{j}}). Similar relations hold for periods around βj\beta_{j}, and for w¯k\bar{w}_{k} derivatives.

The expression for the kinetic energy can in the point vortex case be written, at least formally,

2(ν,ν)1=k,j=1nΓkΓjG(wk,wj)+j=1g(αjηβjηαjηβjη).2(\nu,\nu)_{1}=\sum_{k,j=1}^{n}\Gamma_{k}\Gamma_{j}G(w_{k},w_{j})+\sum_{j=1}^{\texttt{g}}\big{(}\oint_{\alpha_{j}}\eta\oint_{\beta_{j}}*\eta-\oint_{\alpha_{j}}*\eta\oint_{\beta_{j}}\eta\big{)}.

However, the presence of the terms Γk2G(wk,wk)\Gamma_{k}^{2}G(w_{k},w_{k}) makes the first term become infinite. In order to renormalize this singular behavior we isolate the logarithmic pole in the Green function by writing

G(z,w)=12π(log|zw|+H(z,w)),G(z,w)=\frac{1}{2\pi}(-\log|z-w|+H(z,w)), (4.9)

and expand, for a fixed ww, the regular part in a power series in zz as

H(z,w)=h0(w)+12(h1(w)(zw)+h1(w)¯(z¯w¯))+H(z,w)=h_{0}(w)+\frac{1}{2}\left(h_{1}(w)(z-w)+\overline{h_{1}(w)}(\bar{z}-\bar{w})\right)+ (4.10)
+12(h2(w)(zw)2+h2(w)¯(z¯w¯)2)+h11(w)(zw)(z¯w¯)+𝒪(|zw|3).+\frac{1}{2}\left(h_{2}(w)(z-w)^{2}+\overline{h_{2}(w)}(\bar{z}-\bar{w})^{2}\right)+h_{11}(w)(z-w)(\bar{z}-\bar{w})+\mathcal{O}(|z-w|^{3}).

In Appendix 2, Section 11, it is discussed how the coefficients in the expansion (4.10) behave under conformal mapping. For example, the coefficient h11(w)h_{11}(w) transforms as the density of a metric, and it is indeed proportional to the given metric:

h11(w)=π2Vλ(w)2.h_{11}(w)=\frac{\pi}{2V}\lambda(w)^{2}. (4.11)

See (11.5). The coefficients h0(w)h_{0}(w), h1(w)h_{1}(w) and h2(w)h_{2}(w) transform, in certain combinations, as “connections” under conformal mappings. For h0(w)h_{0}(w) this amounts to saying that it too defines a metric, namely the Robin metric, via

ds=eh0(w)|dw|.ds=e^{-h_{0}(w)}|dw|. (4.12)

Metrics of this kind are implicit in the theory of capacity functions, as exposed in [42]. It should be pointed out that our version of the Robin metric depends on the given metric, since the Green function itself depends on it. The Robin metric can also be adapted to various given circulations, and then it becomes more intrinsically hydrodynamic in nature. The coefficient h0(w)h_{0}(w) is one example of a (coordinate) Robin function.

Now, letting ww have the role of being a vortex point indicates that one could renormalize the kinetic energy by simply discarding the singular term log|zw|\log|z-w|, as this seems at first sight to produce just a circular symmetric flow, not affecting the speed of the vortex. However, this is not fully correct in the case of curved surfaces. The term log|zw|\log|z-w| cannot be just removed, it need be replaced by a term which counteracts the inhomogeneous transformation law of h0(w)h_{0}(w) (see (11.8)). Such a term comes naturally from the given metric ds=λ(w)|dw|=elogλ(w)|dw|ds=\lambda(w)|dw|=e^{\log\lambda(w)}|dw|. We see that minus logλ(w)\log\lambda(w) has the right properties, and it combines with h0(w)h_{0}(w) into

Rrobin(w)=12π(h0(w)+logλ(w)).R_{\rm robin}(w)=\frac{1}{2\pi}(h_{0}(w)+\log\lambda(w)). (4.13)

This is indeed a function, the Robin function, and it appears naturally when writing the singularity of the Green function in terms of the distance with respect to the given Riemannian metric:

G(z,w)=12πlogdist(z,w)+Rrobin(w)+𝒪(dist(z,w)).G(z,w)=-\frac{1}{2\pi}\log{\rm dist}(z,w)+R_{\rm robin}(w)+\mathcal{O}({\rm dist}(z,w)).

As for the infinite kinetic energy, we conclude that it should be renormalized by replacing the diagonal terms G(wk,wk)G(w_{k},w_{k}) by Rrobin(wk)R_{\rm robin}(w_{k}). This gives the same equations of motion as more direct approaches, or those available in the literature, like [26, 3, 10]. Thus (twice) the renormalized energy is

2(ν,ν)1,renorm=k=1nΓk2Rrobin(wk)+kjΓkΓjG(wk,wj)+2(\nu,\nu)_{1,{\rm renorm}}=\sum_{k=1}^{n}\Gamma_{k}^{2}R_{\rm robin}(w_{k})+\sum_{k\neq j}\Gamma_{k}\Gamma_{j}G(w_{k},w_{j})+
+j=1g(αjηβjηαjηβjη).+\sum_{j=1}^{\texttt{g}}\big{(}\oint_{\alpha_{j}}\eta\oint_{\beta_{j}}*\eta-\oint_{\alpha_{j}}*\eta\oint_{\beta_{j}}\eta\big{)}. (4.14)

This depends on the locations w1,,wnw_{1},\dots,w_{n} of the vortices (even the last term depends on these, although somewhat more indirectly).

Many authors start out directly with the Robin function (4.13), but there are some advantages with exposing the two terms in it as individual quantities. One is that it clarifies the structure of the vortex motion equations by separating harmonic contributions, such as h0(wk)h_{0}(w_{k}) and G(wk,wj)G(w_{k},w_{j}), from differential geometric contributions, like logλ(wk)\log\lambda(w_{k}). The first category of terms can be classified as nonlocal, coming from solutions of elliptic partial differential equations on the entire surface, whereas the second category are purely local in nature. There is also the contribution from global circulations, represented by the final term in (4.14), and this may be considered to be truly global. So the vortex motion is in principle governed by a balance between different categories of contributions, local, nonlocal, and global in nature. As we shall see later this balance is drastically changed in the more singular case of dipole motion (Sections 7 and 8). Then only the local terms survive.

We need to elaborate further the expression (4.1) and relate it to the global circulations. The 11-form η\eta is harmonic and hence can be expanded as

η=j=1gAjdUβj+j=1gBjdUαj.\eta=-\sum_{j=1}^{\texttt{g}}A_{j}dU_{\beta_{j}}+\sum_{j=1}^{\texttt{g}}B_{j}dU_{\alpha_{j}}. (4.15)

The coefficients AkA_{k}, BkB_{k} are in view of (3.18), (3.19) given by

Ak=αkη,Bk=βkη.A_{k}=\oint_{\alpha_{k}}\eta,\quad B_{k}=\oint_{\beta_{k}}\eta.

The circulations of the total flow ν=ηdGω\nu=\eta-*dG^{\omega} then become

ak=αkν=AkαkdGω,a_{k}=\oint_{\alpha_{k}}\nu=A_{k}-\oint_{\alpha_{k}}*dG^{\omega},
bk=βkν=BkβkdGω.b_{k}=\oint_{\beta_{k}}\nu=B_{k}-\oint_{\beta_{k}}*dG^{\omega}.

We shall treat the circulations a1,,aga_{1},\dots,a_{\texttt{g}}, b1,,bgb_{1},\dots,b_{\texttt{g}} as free (independent) variables, along with the locations w1,,wnw_{1},\dots,w_{n} of the vortices. These variables will be the coordinates of the phase space, and together they determine η\eta and ν\nu. The locations of the vortices are complex variables, in contrast to the circulations which are real.

If γ\gamma is a closed oriented curve in MM, fixed in time and avoiding the point vortices, then according to the Euler equation (2.13), and in the notation used there, the circulation of ν\nu around γ\gamma changes with speed

ddtγν=γνt=γ(dϕ𝐯ν)=γi(𝐯)𝑑ν=\frac{d}{dt}\oint_{\gamma}\nu=\oint_{\gamma}\frac{\partial\nu}{\partial t}=\oint_{\gamma}(d\phi-\mathcal{L}_{\bf v}\nu)=-\oint_{\gamma}i({\bf v})d\nu=
=γi(𝐯)ω=γi(𝐯)(ΓVvol)=ΓVγν.=-\oint_{\gamma}i({\bf v})\omega=\oint_{\gamma}i({\bf v})\big{(}\frac{\Gamma}{V}{\rm vol}\big{)}=\frac{\Gamma}{V}\oint_{\gamma}*\nu.

Here we used also (2.7), (2.8), (4.4). It follows in particular, on choosing γ=αk,βk\gamma=\alpha_{k},\beta_{k}, that

dakdt=ΓVαkν=ΓVαkη,\frac{da_{k}}{dt}=\frac{\Gamma}{V}\oint_{\alpha_{k}}*\nu=\frac{\Gamma}{V}\oint_{\alpha_{k}}*\eta, (4.16)
dbkdt=ΓVβkν=ΓVβkη.\frac{db_{k}}{dt}=\frac{\Gamma}{V}\oint_{\beta_{k}}*\nu=\frac{\Gamma}{V}\oint_{\beta_{k}}*\eta. (4.17)

4.2 Matrix formalism and the Hamiltonian

The period matrix (written here in block form)

(PRRTQ)=((βkdUβj)(βkdUαj)(αkdUβj)(αkdUαj))=\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)=\left(\begin{array}[]{cc}(-\oint_{\beta_{k}}*dU_{\beta_{j}})&(\oint_{\beta_{k}}*dU_{\alpha_{j}})\\ (\oint_{\alpha_{k}}*dU_{\beta_{j}})&(-\oint_{\alpha_{k}}*dU_{\alpha_{j}})\\ \end{array}\right)=
=((MdUβkdUβj)(MdUβkdUαj)(MdUαkdUβj)(MdUαkdUαj))=\left(\begin{array}[]{cc}(\int_{M}dU_{\beta_{k}}\wedge*dU_{\beta_{j}})&(-\int_{M}dU_{\beta_{k}}\wedge*dU_{\alpha_{j}})\\ (-\int_{M}dU_{\alpha_{k}}\wedge*dU_{\beta_{j}})&(\int_{M}dU_{\alpha_{k}}\wedge*dU_{\alpha_{j}})\\ \end{array}\right) (4.18)

is symmetric and positive definite (see [12] in general). In particular, PP and QQ are themselves symmetric and positive definite. As for RR and RTR^{T} we need to be explicit with what are row and column indices above: in all entries above, kk is the row index, jj the column index. Thus, for example, Rkj=βkdUαjR_{kj}=\oint_{\beta_{k}}*dU_{\alpha_{j}}.

Next we write

dUα=(dUα1dUαg),dUβ=(dUβ1dUβg).dU_{\alpha}=\left(\begin{array}[]{c}dU_{\alpha_{1}}\\ \vdots\\ dU_{\alpha_{\texttt{g}}}\end{array}\right),\quad dU_{\beta}=\left(\begin{array}[]{c}dU_{\beta_{1}}\\ \vdots\\ dU_{\beta_{\texttt{g}}}\end{array}\right).

As mentioned, these two column matrices together define a basis of the harmonic forms. Another basis is provided by the corresponding Hodge starred column vectors dUα*dU_{\alpha}, dUβ*dU_{\beta}, in similar matrix notation. The relation between the bases is

Lemma 4.1.

The two bases {dUα,dUβ}\{dU_{\alpha},dU_{\beta}\} and {dUα,dUβ}\{*dU_{\alpha},*dU_{\beta}\} are related by

(dUαdUβ)=(RTQPR)(dUαdUβ).\left(\begin{array}[]{cc}*dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right)=\left(\begin{array}[]{cc}R^{T}&Q\\ -P&-R\\ \end{array}\right)\left(\begin{array}[]{cc}dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right). (4.19)
Proof.

One simply checks that the two members in (4.19) have the same periods with respect to the homology basis {αj,βj}\{\alpha_{j},\beta_{j}\}. ∎

We arrange also the circulations of the flow ν\nu into column vectors:

a=(a1ag)=(α1ναgν),b=(b1bg)=(β1νβgν),a=\left(\begin{array}[]{c}a_{1}\\ \vdots\\ a_{\texttt{g}}\end{array}\right)=\left(\begin{array}[]{c}\oint_{\alpha_{1}}\nu\\ \vdots\\ \oint_{\alpha_{\texttt{g}}}\nu\end{array}\right),\quad b=\left(\begin{array}[]{c}b_{1}\\ \vdots\\ b_{\texttt{g}}\end{array}\right)=\left(\begin{array}[]{c}\oint_{\beta_{1}}\nu\\ \vdots\\ \oint_{\beta_{\texttt{g}}}\nu\end{array}\right),

briefly written as

a=αν,b=βν.a=\oint_{\alpha}\nu,\quad b=\oint_{\beta}\nu. (4.20)

Similarly for other vectors of circulations, for example

A=αη=αν+αdGω=a+αdGω,A=\oint_{\alpha}\eta=\oint_{\alpha}\nu+\oint_{\alpha}*dG^{\omega}=a+\oint_{\alpha}*dG^{\omega}, (4.21)
B=βη=βν+βdGω=b+βdGω.B=\oint_{\beta}\eta=\oint_{\beta}\nu+\oint_{\beta}*dG^{\omega}=b+\oint_{\beta}*dG^{\omega}. (4.22)

The α\alpha-periods of the conjugated Green function were computed in (4.6), and from (4.7), (4.8) it follows that

Awkdwk+Aw¯kdw¯k=ΓkdUα(wk).\frac{\partial A}{\partial w_{k}}dw_{k}+\frac{\partial A}{\partial\bar{w}_{k}}d\bar{w}_{k}={\Gamma_{k}}\,dU_{\alpha}(w_{k}).

Taking into account also the dependence of a1,,aga_{1},\dots,a_{\texttt{g}}, and doing the same for BB and the β\beta-periods, gives the total differentials

dA=da+dαdGω=da+k=1nΓkdUα(wk),dA=da+d\oint_{\alpha}*dG^{\omega}=da+\sum_{k=1}^{n}\Gamma_{k}dU_{\alpha}(w_{k}), (4.23)
dB=db+dβdGω=db+k=1nΓkdUβ(wk).dB=db+d\oint_{\beta}*dG^{\omega}=db+\sum_{k=1}^{n}\Gamma_{k}dU_{\beta}(w_{k}). (4.24)

In matrix notation (4.15) becomes

η=ATdUβ+BTdUα=(BT,AT)(dUαdUβ),\eta=-A^{T}dU_{\beta}+B^{T}dU_{\alpha}=\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{c}dU_{\alpha}\\ dU_{\beta}\end{array}\right), (4.25)

and the contribution from η\eta to the kinetic energy is the quadratic form

Mηη=k=1g(αkηβkηβkηαkη)=\int_{M}\eta\wedge*\eta=\sum_{k=1}^{\texttt{g}}(\oint_{\alpha_{k}}\eta\oint_{\beta_{k}}*\eta-\oint_{\beta_{k}}\eta\oint_{\alpha_{k}}*\eta)=
=(AT,BT)(PRRTQ)(AB).=\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right).

The last equality is based on straight-forward computations. Note that PP, QQ, RR are fixed matrices, while AA and BB depend on w1,,wnw_{1},\dots,w_{n}, a1,,aga_{1},\dots,a_{\texttt{g}}, b1,,bgb_{1},\dots,b_{\texttt{g}} via (4.21), (4.22).

We now define the Hamiltonian function, \mathcal{H}, as the renormalized kinetic energy of the flow considered as a function of the locations of the point vortices and of the circulations:

2(w1,,wn;a1,,ag,b1,,bg)=2(ν,ν)1,renorm=2\mathcal{H}(w_{1},\dots,w_{n};a_{1},\dots,a_{\texttt{g}},b_{1},\dots,b_{\texttt{g}})=2(\nu,\nu)_{1,{\rm renorm}}= (4.26)
(Γ1,Γ2Γn)(Rrobin(w1)G(w1,w2)G(w1,wn)G(w2,w1)Rrobin(w2)G(w2,wn)G(wn,w1)G(wn,w2)Rrobin(wn))(Γ1Γ2Γn)+\left(\begin{array}[]{cccc}\Gamma_{1},&\Gamma_{2}&\ldots&\Gamma_{n}\end{array}\right)\left(\begin{array}[]{cccc}R_{\rm robin}(w_{1})&G(w_{1},w_{2})&\ldots&G(w_{1},w_{n})\\ G(w_{2},w_{1})&R_{\rm robin}(w_{2})&\ldots&G(w_{2},w_{n})\\ \vdots&\vdots&\ddots&\vdots\\ G(w_{n},w_{1})&G(w_{n},w_{2})&\dots&R_{\rm robin}(w_{n})\end{array}\right)\left(\begin{array}[]{c}\Gamma_{1}\\ \Gamma_{2}\\ \vdots\\ \Gamma_{n}\end{array}\right)+
+(AT,BT)(PRRTQ)(AB).+\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right).

More formally one could consider the complex conjugates w¯j\bar{w}_{j} of the vortex positions as independent variables and write the Hamiltonian as

(w1,,wn,w¯1,,w¯n;a1,,ag,b1,,bg).\mathcal{H}(w_{1},\dots,w_{n},\bar{w}_{1},\dots,\bar{w}_{n};a_{1},\dots,a_{\texttt{g}},b_{1},\dots,b_{\texttt{g}}).

5 Hamilton’s equations

5.1 Phase space and symplectic form

To formulate Hamilton’s equation one needs, besides the Hamiltonian function itself, a phase space and a symplectic form on it. The phase space will in our case consist of all possible configurations of the vortices, collisions not allowed, together with all possible circulations of the flow around the curves in the homology basis. Thus we take it to be

={(w1,,wn;a1,,ag,b1,,bg):wjM,wkwj for kj}.\mathcal{F}=\{(w_{1},\dots,w_{n};a_{1},\dots,a_{\texttt{g}},b_{1},\dots,b_{\texttt{g}}):w_{j}\in M,w_{k}\neq w_{j}\text{ for }k\neq j\}.

Here the wjw_{j} shall be interpreted as points on MM, but in most equations below they will refer to complex coordinates for such points. Compare Remark 3.2.

Assuming that Γ0\Gamma\neq 0 (recall (4.3)) the symplectic form on \mathcal{F} is taken to be

Ω=k=1nΓkvol(wk)VΓj=1gdajdbj=\Omega=\sum_{k=1}^{n}\Gamma_{k}{\rm vol}(w_{k})-\frac{V}{\Gamma}\sum_{j=1}^{\texttt{g}}da_{j}\wedge db_{j}=
=12ik=1nΓkλ(wk)2dwkdw¯kΓVj=1gdajΓdbjΓ.=-\frac{1}{2\mathrm{i}}\sum_{k=1}^{n}\Gamma_{k}\lambda(w_{k})^{2}dw_{k}\wedge d\bar{w}_{k}-\Gamma{V}\sum_{j=1}^{\texttt{g}}\frac{da_{j}}{\Gamma}\wedge\frac{db_{j}}{\Gamma}. (5.1)

The last expression makes sense also if Γ=0\Gamma=0, because by (4.16), (4.17) the factors daj/Γda_{j}/\Gamma and dbj/Γdb_{j}/\Gamma remain finite as Γ0\Gamma\to 0. Thus in both expressions above, the last term shall simply be removed if Γ=0\Gamma=0.

Let

ξ=k=1n(w˙kwk+w¯˙kw¯k)+j=1g(a˙jaj+b˙jbj)\xi=\sum_{k=1}^{n}(\dot{w}_{k}\frac{\partial}{\partial w_{k}}+\dot{\bar{w}}_{k}\frac{\partial}{\partial\bar{w}_{k}})+\sum_{j=1}^{\texttt{g}}(\dot{a}_{j}\frac{\partial}{\partial a_{j}}+\dot{b}_{j}\frac{\partial}{\partial b_{j}})

denote a generic tangent vector of \mathcal{F} viewed as a derivation. As for the first term, recall the conventions in Remark 3.3. The Hamilton equations in general say that

i(ξ)Ω=d.i(\xi)\Omega=d\mathcal{H}. (5.2)

One main issue then is to verify that with our choices of phase space, symplectic form and Hamiltonian function, the equations (5.2) really produce the expected vortex dynamics. This will be accomplished in Section 5.2.

To evaluate (5.2) we first compute the left member as

i(ξ)Ω=12ik=1nΓkλ(wk)2(w˙kdw¯kw¯˙kdwk)VΓj=1g(a˙jdbjb˙jdaj).i(\xi)\Omega=-\frac{1}{2\mathrm{i}}\sum_{k=1}^{n}\Gamma_{k}\lambda(w_{k})^{2}\big{(}\dot{w}_{k}d\bar{w}_{k}-\dot{\bar{w}}_{k}d{w}_{k}\big{)}-\frac{V}{\Gamma}\sum_{j=1}^{\texttt{g}}(\dot{a}_{j}db_{j}-\dot{b}_{j}da_{j}). (5.3)

Explicitly (5.2) therefore says that

Γkλ(wk)2w˙k=2iw¯k,\Gamma_{k}\lambda(w_{k})^{2}{\dot{w}_{k}}=-2\mathrm{i}\frac{\partial\mathcal{H}}{\partial\bar{w}_{k}}, (5.4)
a˙j=ΓVbj,b˙j=+ΓVaj.\dot{a}_{j}=-\frac{\Gamma}{V}\frac{\partial\mathcal{H}}{\partial{b}_{j}},\quad\dot{b}_{j}=+\frac{\Gamma}{V}\frac{\partial\mathcal{H}}{\partial{a}_{j}}. (5.5)

Here the partial derivatives in the right members are implicit in

d=k=1nΓk22(Rrobin(wk)wkdwk+Rrobin(wk)w¯kdw¯k)+d\mathcal{H}=\sum_{k=1}^{n}\frac{\Gamma_{k}^{2}}{2}\big{(}\frac{\partial R_{\rm robin}(w_{k})}{\partial w_{k}}dw_{k}+\frac{\partial R_{\rm robin}(w_{k})}{\partial\bar{w}_{k}}d\bar{w}_{k}\big{)}+ (5.6)
+k,j=1,kjnΓkΓj2(G(wk,wj)wkdwk+G(wk,wj)w¯kdw¯k)++\sum_{k,j=1,k\neq j}^{n}\frac{\Gamma_{k}\Gamma_{j}}{2}\big{(}\frac{\partial G(w_{k},w_{j})}{\partial{w}_{k}}d{w}_{k}+\frac{\partial G(w_{k},w_{j})}{\partial\bar{w}_{k}}d\bar{w}_{k}\big{)}+
+k,j=1,kjnΓkΓj2(G(wk,wj)wjdwj+G(wk,wj)w¯jdw¯j)++\sum_{k,j=1,k\neq j}^{n}\frac{\Gamma_{k}\Gamma_{j}}{2}\big{(}\frac{\partial G(w_{k},w_{j})}{\partial{w}_{j}}d{w}_{j}+\frac{\partial G(w_{k},w_{j})}{\partial\bar{w}_{j}}d\bar{w}_{j}\big{)}+
+(AT,BT)(PRRTQ)(k=1nΓkdUα(wk)k=1nΓkdUβ(wk))++\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}\sum_{k=1}^{n}\Gamma_{k}dU_{\alpha}(w_{k})\\ \\ \sum_{k=1}^{n}\Gamma_{k}dU_{\beta}(w_{k})\end{array}\right)+
+(AT,BT)(PRRTQ)(dadb).+\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}da\\ db\end{array}\right).

Recall here the expressions (4.23), (4.24) for AA, BB in terms of the phase space variables aa, bb (as column matrices) and w1,,wnw_{1},\dots,w_{n}, w¯1,,w¯n\bar{w}_{1},\dots,\bar{w}_{n}.

In the partial derivatives of RrobinR_{\rm robin} we single out the two affine connections

rmetric(w)=2wlogλ(w),r_{\rm metric}(w)=2\frac{\partial}{\partial w}\log\lambda(w),\qquad\qquad (5.7)
rrobin(w)=2h1(w)=2wh0(w),r_{\rm robin}(w)\ =-2h_{1}(w)=-2\frac{\partial}{\partial w}h_{0}(w), (5.8)

(see (11.3) for the last equality). Recalling (4.13) we then have the following alternative expressions for (ingredients of) the first term in dd\mathcal{H}:

Rrobin(wk)wk=14π(rmetric(wk)rrobin(wk))=\frac{\partial R_{\rm robin}(w_{k})}{\partial w_{k}}=\frac{1}{4\pi}\big{(}r_{\rm metric}(w_{k})-r_{\rm robin}(w_{k})\big{)}= (5.9)
=12π(h1(wk)+wklogλ(wk)).=\frac{1}{2\pi}\big{(}{h_{1}(w_{k})}+\frac{\partial}{\partial{w}_{k}}\log\lambda(w_{k})\big{)}.

Similarly for the conjugated quantities.

5.2 Dynamical equations

The following theorem now makes the Hamilton equations (5.2) explicit in the vortex dynamics case.

Theorem 5.1.

The dynamical equations for the vortices wkw_{k} and the circulations aja_{j}, bjb_{j} are, in matrix notation,

λ(wk)2w˙k=\displaystyle\lambda(w_{k})^{2}{\dot{w}_{k}}= Γk2πi(h1(wk)¯+w¯klogλ(wk))2ij=1,jknΓjG(wk,wj)w¯k\displaystyle\frac{\Gamma_{k}}{2\pi\mathrm{i}}\big{(}\overline{h_{1}(w_{k})}+\frac{\partial}{\partial\bar{w}_{k}}\log\lambda(w_{k})\big{)}-{2\mathrm{i}}\sum_{j=1,j\neq k}^{n}\Gamma_{j}\frac{\partial G(w_{k},w_{j})}{\partial\bar{w}_{k}}-
+2(BT,AT)(Uα(wk)/w¯kUβ(wk)/w¯k),\displaystyle+2\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{c}{\partial U_{\alpha}(w_{k})}/{\partial\bar{w}_{k}}\\ {\partial U_{\beta}(w_{k})}/{\partial\bar{w}_{k}}\end{array}\right),
(a˙b˙)=\displaystyle\left(\begin{array}[]{c}\dot{a}\\ \dot{b}\end{array}\right)= ΓV(RTQPR)(AB).\displaystyle\frac{\Gamma}{V}\left(\begin{array}[]{cc}-R^{T}&-Q\\ P&R\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right).

Recall the column vectors (4.21), (4.22):

A=a+αdGω,B=b+βdGω.A=a+\oint_{\alpha}*dG^{\omega},\quad B=b+\oint_{\beta}*dG^{\omega}.
Proof.

The equations are obtained by identifying the expression (5.3) for i(ξ)Ωi(\xi)\Omega with the expression (5.6) for dd\mathcal{H}. Thus, in (5.3), the kk:th term in the first sum,

12iΓkλ(wk)2(w˙kdw¯kw¯˙kdwk),-\frac{1}{2\mathrm{i}}\Gamma_{k}\lambda(w_{k})^{2}\big{(}\dot{w}_{k}d\bar{w}_{k}-\dot{\bar{w}}_{k}d{w}_{k}\big{)}, (5.10)

is to be identified with the corresponding parts in the right member of (5.6), namely the first three terms. Together with (5.9) this gives immediately the first two terms in the equation for w˙k\dot{w}_{k}.

The third term comes from the term

(AT,BT)(PRRTQ)(k=1nΓkdUα(wk)k=1nΓkdUβ(wk))\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}\sum_{k=1}^{n}\Gamma_{k}dU_{\alpha}(w_{k})\\ \\ \sum_{k=1}^{n}\Gamma_{k}dU_{\beta}(w_{k})\end{array}\right)

in equation (5.6). On using (4.19) this can be rewritten as

(BT,AT)(RTQPR)(k=1nΓkdUα(wk)k=1nΓkdUβ(wk))=\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{cc}R^{T}&Q\\ -P&-R\\ \end{array}\right)\left(\begin{array}[]{c}\sum_{k=1}^{n}\Gamma_{k}dU_{\alpha}(w_{k})\\ \\ \sum_{k=1}^{n}\Gamma_{k}dU_{\beta}(w_{k})\end{array}\right)=
=(BT,AT)(k=1nΓkdUα(wk)k=1nΓkdUβ(wk)).=\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{c}\sum_{k=1}^{n}\Gamma_{k}*dU_{\alpha}(w_{k})\\ \\ \sum_{k=1}^{n}\Gamma_{k}*dU_{\beta}(w_{k})\end{array}\right).

Identifying here the coefficient for dw¯kd\bar{w}_{k} with the corresponding coefficient in (5.10) gives

λ(wk)2w˙k=2(BT,AT)(Uα(wk)w¯kUβ(wk)w¯k),\lambda(w_{k})^{2}\dot{w}_{k}=2\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{c}\frac{\partial U_{\alpha}(w_{k})}{\partial\bar{w}_{k}}\\ \frac{\partial U_{\beta}(w_{k})}{\partial\bar{w}_{k}}\end{array}\right),

as desired.

Finally, the term

(AT,BT)(PRRTQ)(dadb).\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}da\\ db\end{array}\right).

in (5.6) is to be identified with

VΓ(b˙,a˙)(dadb)-\frac{V}{\Gamma}\left(\begin{array}[]{cc}-\dot{b},&\dot{a}\\ \end{array}\right)\left(\begin{array}[]{cc}da\\ db\\ \end{array}\right)

in (5.3). This gives

(a˙b˙)=ΓV(RTQPR)(AB),\left(\begin{array}[]{c}\dot{a}\\ \dot{b}\end{array}\right)=\frac{\Gamma}{V}\left(\begin{array}[]{cc}-R^{T}&-Q\\ P&R\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right),

as desired. ∎

Besides the formal proof of the theorem it is essential to show that the dynamics given in Theorem 5.1 really is the “expected dynamics”. For the vortices this expected dynamics is that each individual vortex, say wkw_{k}, moves with the speed of total flow, namely the 11-form ν=ηdGω\nu=\eta-*dG^{\omega} (see (4.1)) converted into a vector, however with the modification that the contribution to ν\nu from wkw_{k} itself shall be regularized according to standard procedures involving the Robin function.

In the notation of Remark 3.3, the velocity 𝐕{\bf V} of the kk:th vortex wkw_{k} corresponds, with z=wkz=w_{k}, to the covector λ22(z¯˙dz+z˙dz¯)\frac{\lambda^{2}}{2}(\dot{\bar{z}}dz+\dot{{z}}d\bar{z}), see equations (3.11), (3.12). Thus in the notation of the Theorem 5.1 the expected dynamics is expressed by the equation

η{dGω}renormalized=12λ(wk)2(w¯˙kdwk+w˙kdw¯k),\eta-\{*dG^{\omega}\}_{\rm renormalized}=\frac{1}{2}\lambda(w_{k})^{2}\big{(}\dot{\bar{w}}_{k}dw_{k}+{\dot{w}_{k}}d\bar{w}_{k}\big{)},

the left member to be evaluated at wkw_{k}. Here the second term in the left member is the well-known [37, 3] contribution from the Green function, where (referring to (5.9), (4.5))

{dGω}renormalized(z)=ΓkdRrobin(z)+j=1,jknΓjdG(z,wj),\{*dG^{\omega}\}_{\rm renormalized}(z)=\Gamma_{k}*dR_{\rm robin}(z)+\sum_{j=1,j\neq k}^{n}\Gamma_{j}*dG(z,w_{j}),

evaluated at z=wkz=w_{k} and the Hodge star taken with respect to zz. This gives

{dGω}renormalized(wk)=iΓk(Rrobin(wk)wkdwkRrobin(wk)w¯kdw¯k)+-\{*dG^{\omega}\}_{\rm renormalized}(w_{k})=\mathrm{i}\Gamma_{k}\Big{(}\frac{\partial R_{\rm robin}(w_{k})}{\partial w_{k}}dw_{k}-\frac{\partial R_{\rm robin}(w_{k})}{\partial\bar{w}_{k}}d\bar{w}_{k}\Big{)}+
+ij=1,jknΓj(G(wk,wj)wkdwkG(wk,wj)w¯kdw¯k)+{\mathrm{i}}\sum_{j=1,j\neq k}^{n}\Gamma_{j}\Big{(}\frac{\partial G(w_{k},w_{j})}{\partial{w}_{k}}dw_{k}-\frac{\partial G(w_{k},w_{j})}{\partial\bar{w}_{k}}d\bar{w}_{k}\Big{)}

The above accounts for the first two terms in the right member of Theorem 5.1. The third term is the contribution from η\eta, and can by (4.25) be directly identified as 2ηz¯2\eta_{\bar{z}} evaluated at z=wkz=w_{k}.

Finally we verify that the periods change according to (4.16), (4.17). By (4.25), (4.19) we have

η=(BT,AT)(dUαdUβ)=(BT,AT)(RTQPR)(dUαdUβ)=*\eta=\left(\begin{array}[]{cc}B^{T},&-A^{T}\\ \end{array}\right)\left(\begin{array}[]{cc}*dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right)=\left(\begin{array}[]{cc}B^{T},&-A^{T}\\ \end{array}\right)\left(\begin{array}[]{cc}R^{T}&Q\\ -P&-R\\ \end{array}\right)\left(\begin{array}[]{cc}dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right)=
=((dUα)T,(dUβ)T)(PRRTQ)(AB).=\left(\begin{array}[]{cc}(dU_{\alpha})^{T},&(dU_{\beta})^{T}\\ \end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{cc}A\\ B\\ \end{array}\right).

By (3.18), (3.19) integration of this gives

(αηβη)=(RTQPR)(AB).\left(\begin{array}[]{cc}\oint_{\alpha}*\eta\\ \oint_{\beta}*\eta\\ \end{array}\right)=\left(\begin{array}[]{cc}-R^{T}&-Q\\ P&R\end{array}\right)\left(\begin{array}[]{cc}A\\ B\\ \end{array}\right).

From this we see that the last equation in the theorem is in exact agreement with (4.16), (4.17), as claimed.

6 Motion of a single point vortex

In the case of a single vortex Theorem 5.1 simplifies a little. We may then denote the vortex point w1w_{1} simply by ww, and the strength Γ1\Gamma_{1} agrees with the total vorticity Γ\Gamma for the point vortices. If in addition g=0\texttt{g}=0 then everything simplifies considerable. There is only one free variable, the location wMw\in M for the vortex, and the dynamical equation for this is

λ(w)2dwdt=Γ2πi(h1(w)¯+w¯logλ(w)).\lambda(w)^{2}\frac{dw}{dt}=\frac{\Gamma}{2\pi\mathrm{i}}\big{(}\overline{h_{1}(w)}+\frac{\partial}{\partial\bar{w}}\log\lambda(w)\big{)}.

The Hamiltonian and the symplectic form are

(w)=Γ2Rrobin(w)=Γ22π(h0(w)+logλ(w)),\mathcal{H}(w)=\Gamma^{2}R_{\rm robin}(w)=\frac{\Gamma^{2}}{2\pi}(h_{0}(w)+\log\lambda(w)),
Ω=Γvol(w)=12iΓλ(w)2dwdw¯.\Omega=\Gamma\,{\rm vol}(w)=-\frac{1}{2\mathrm{i}}\Gamma\lambda(w)^{2}dw\wedge d\bar{w}.

It follows that if (and only if) the two metrics

dsmetric2=λ(w)2|dw|2=2Vπh11(w)|dw|2,ds_{\rm metric}^{2}=\lambda(w)^{2}|dw|^{2}=\frac{2V}{\pi}h_{11}(w)|dw|^{2}, (6.1)
dsrobin2=e2h0(w)|dw|2ds_{\rm robin}^{2}\ =e^{-2h_{0}(w)}|dw|^{2}\qquad\qquad\qquad\quad (6.2)

are identical, up to a constant factor, then the vortex will never move, whatever its initial position is. In [3, 19] this is referred to as dsmetricds_{\rm metric} being a “steady vortex metric”, or being “hydrodynamically neutral”.

Example 6.1.

An obvious example is a homogenous sphere. Indeed, for a sphere of radius one we have, in coordinates obtained by stereographic projection into the complex plane, well-known formulas such as

G(z,w)=14π(log|zw|2(1+|z|2)(1+|w|2)+1),G(z,w)=-\frac{1}{4\pi}\Big{(}\log\frac{|z-w|^{2}}{(1+|z|^{2})(1+|w|^{2})}+1\Big{)},
h0(w)=log(1+|w|2)12,h1(w)=w¯1+|w|2,h2(w)=w¯22(1+|w|2)2,h_{0}(w)=\log(1+|w|^{2})-\frac{1}{2},\quad h_{1}(w)=\frac{\bar{w}}{1+|w|^{2}},\quad h_{2}(w)=\frac{\bar{w}^{2}}{2(1+|w|^{2})^{2}},
e2h0(w)=e(1+|w|2)2,λ(w)2=8h11(w)=4(1+|w|2)2.e^{-2h_{0}(w)}=\frac{e}{(1+|w|^{2})^{2}},\quad\lambda(w)^{2}=8h_{11}(w)=\frac{4}{(1+|w|^{2})^{2}}.

The Hamiltonian function is constant,

(w)=Γ22π(log(1+|w|2)12+log21+|w|2)=Γ24π(2log21),\mathcal{H}(w)=\frac{\Gamma^{2}}{2\pi}\big{(}\log(1+|w|^{2})-\frac{1}{2}+\log\frac{2}{1+|w|^{2}}\big{)}=\frac{\Gamma^{2}}{4\pi}(2\log 2-1),

and there is no motion of the vortex.

When g>0\texttt{g}>0 one has to include also the circulating flows in the picture, so that the dynamical equations become, by Theorem 5.1,

λ(w)2w˙=\displaystyle\lambda(w)^{2}\dot{w}= Γ2πi(h1(w)¯+w¯logλ(w))+2(BT,AT)(Uα(w)/w¯Uβ(w)/w¯),\displaystyle\frac{\Gamma}{2\pi\mathrm{i}}\big{(}\overline{h_{1}(w)}+\frac{\partial}{\partial\bar{w}}\log\lambda(w)\big{)}+2\left(\begin{array}[]{cc}B^{T},&-A^{T}\end{array}\right)\left(\begin{array}[]{c}{\partial U_{\alpha}(w)}/{\partial\bar{w}}\\ {\partial U_{\beta}(w)}/{\partial\bar{w}}\end{array}\right),
(a˙b˙)=\displaystyle\left(\begin{array}[]{c}\dot{a}\\ \dot{b}\end{array}\right)= ΓV(RTQPR)(AB).\displaystyle\frac{\Gamma}{V}\left(\begin{array}[]{cc}-R^{T}&-Q\\ P&R\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right).

7 Motion of a vortex pair in the dipole limit

For a vortex pair {w1,w2}\{w_{1},w_{2}\} with Γ1=Γ2\Gamma_{1}=-\Gamma_{2} we have Γ=Γ1+Γ2=0\Gamma=\Gamma_{1}+\Gamma_{2}=0, hence there is no compensating background vorticity. The circulations aa and bb will be time independent by the last equation in Theorem 5.1 and are not needed in phase space, which then simply becomes

={(w1,w2):w1,w2M,w1w2},\mathcal{F}=\{(w_{1},w_{2}):w_{1},w_{2}\in M,w_{1}\neq w_{2}\},

with symplectic form

Ω=12iΓ1(λ(w1)2dw1dw¯1λ(w2)2dw2dw¯2).\Omega=-\frac{1}{2\mathrm{i}}\Gamma_{1}\big{(}\lambda(w_{1})^{2}dw_{1}\wedge d\bar{w}_{1}-\lambda(w_{2})^{2}dw_{2}\wedge d\bar{w}_{2}\big{)}. (7.1)

The Hamiltonian is the same quantity as before, see (4.26), but it may now be considered as a function only of w1w_{1} and w2w_{2}. The circulations aa, bb are fixed parameters, given in advance.

However the period vectors AA and BB still depend on time via w1w_{1}, w2w_{2}. Indeed, using (3.20), (3.21) we have

A=a+Γ1α(dG(,w1)dG(,w2))=a+Γ1w2w1dUα,A=a+\Gamma_{1}\oint_{\alpha}\big{(}*dG(\cdot,w_{1})-*dG(\cdot,w_{2})\big{)}=a+\Gamma_{1}\int_{w_{2}}^{w_{1}}dU_{\alpha},
B=b+Γ1β(dG(,w1)dG(,w2))=b+Γ1w2w1dUβj.B=b+\Gamma_{1}\oint_{\beta}\big{(}*dG(\cdot,w_{1})-*dG(\cdot,w_{2})\big{)}=b+\Gamma_{1}\int_{w_{2}}^{w_{1}}dU_{\beta_{j}}.

The Hamiltonian is

2(w1,w2)=(Γ1,Γ1)(Rrobin(w1)G(w1,w2)G(w2,w1)Rrobin(w2))(Γ1Γ1)+2\mathcal{H}(w_{1},w_{2})=\left(\begin{array}[]{cc}\Gamma_{1},&-\Gamma_{1}\end{array}\right)\left(\begin{array}[]{cc}R_{\rm robin}(w_{1})&G(w_{1},w_{2})\\ G(w_{2},w_{1})&R_{\rm robin}(w_{2})\\ \end{array}\right)\left(\begin{array}[]{c}\Gamma_{1}\\ -\Gamma_{1}\end{array}\right)+
+(AT,BT)(PRRTQ)(AB),+\left(\begin{array}[]{cc}A^{T},&B^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}A\\ B\end{array}\right), (7.2)

and the dynamics of the vortex pair {w1,w2}\{w_{1},w_{2}\} becomes, by (5.4),

Γ1λ(w1)2w˙1=2i(w1,w2)w¯1,\Gamma_{1}\lambda(w_{1})^{2}\,\dot{w}_{1}=-2\mathrm{i}\,\frac{\partial\mathcal{H}(w_{1},w_{2})}{\partial\bar{w}_{1}},
Γ1λ(w2)2w˙2=+2i(w1,w2)w¯2.\Gamma_{1}\lambda(w_{2})^{2}\,\dot{w}_{2}=+2\mathrm{i}\,\frac{\partial\mathcal{H}(w_{1},w_{2})}{\partial\bar{w}_{2}}.

In place of w1w_{1} and w2w_{2} we may turn to w=12(w1+w2)w=\frac{1}{2}(w_{1}+w_{2}) and u=12(w1w2)u=\frac{1}{2}(w_{1}-w_{2}) as coordinates. These are similar to the “center-arrow coordinates” used in [3, 38]. Then

{w1=w+u,w2=wu.\begin{cases}w_{1}=w+u,\\ w_{2}=w-u.\end{cases} (7.3)

We are interested in the limit u0u\to 0, and Taylor expansion of H(w1,w2)=H(w+u,wu)H(w_{1},w_{2})=H(w+u,w-u) with respect to uu, u¯\bar{u} gives, using relations in Appendix 2,

H(w+u,wu)=h0(w)Re((4h2(w)h1(w)w)u2)+H(w+u,w-u)=h_{0}(w)-\operatorname{Re}\big{(}(4h_{2}(w)-\frac{\partial h_{1}(w)}{\partial w})u^{2}\big{)}+
+(4h11(w)h1(w)w¯)|u|2+𝒪(|u|3).+(4h_{11}(w)-\frac{\partial h_{1}(w)}{\partial\bar{w}})|u|^{2}+\mathcal{O}(|u|^{3}). (7.4)

The linear terms vanish because of the symmetry of H(w1,w2)H(w_{1},w_{2}). The second order terms will only be used briefly in Section 8, and even the constant term h0(w)h_{0}(w) will eventually disappear below. In addition to (7.4) we have the Taylor expansions

h0(w±u)=h0(w)±(h1(w)u+h1(w)u¯)+𝒪(|u|2),h_{0}(w\pm u)=h_{0}(w)\pm\big{(}h_{1}(w)u+\overline{{h}_{1}(w)u}\big{)}+\mathcal{O}(|u|^{2}),
logλ(w±u)=logλ(w)±12(r(w)u+r(w)u¯)+𝒪(|u|2)\log\lambda(w\pm u)=\log\lambda(w)\pm\frac{1}{2}\big{(}r(w)u+\overline{r(w)u}\big{)}+\mathcal{O}(|u|^{2})

(coupled signs throughout). The latter equation uses the affine connection r(w)=rmetric(w)r(w)=r_{\rm metric}(w), see (5.7) or (10.1). For later use we record also the expansion

λ(w±u)2=λ(w)2(1±(r(w)u+r(w)¯u¯)+𝒪(|u|2)).\lambda(w\pm u)^{2}=\lambda(w)^{2}\Big{(}1\pm\big{(}r(w)u+\overline{r(w)}\bar{u}\big{)}+\mathcal{O}(|u|^{2})\Big{)}. (7.5)

Using these expansions we obtain, for the first matrix in the Hamiltonian (7.2),

2π(Rrobin(w1)G(w1,w2)G(w2,w1)Rrobin(w2))=2\pi\left(\begin{array}[]{cc}R_{\rm robin}(w_{1})&G(w_{1},w_{2})\\ G(w_{2},w_{1})&R_{\rm robin}(w_{2})\\ \end{array}\right)= (7.6)
=(h0(w+u)+logλ(w+u)log|2u|+H(w+u,wu)log|2u|+H(w+u,wu)h0(wu)+logλ(wu))==\left(\begin{array}[]{cc}h_{0}(w+u)+\log\lambda(w+u)&-\log|2u|+H(w+u,w-u)\\ -\log|2u|+H(w+u,w-u)&h_{0}(w-u)+\log\lambda(w-u)\\ \end{array}\right)=
=(logλ(w)log|2u|log|2u|logλ(w))+(h0(w)h0(w)h0(w)h0(w))+=\left(\begin{array}[]{cc}\log\lambda(w)&-\log|2u|\\ -\log|2u|&\log\lambda(w)\\ \end{array}\right)+\left(\begin{array}[]{cc}h_{0}(w)&h_{0}(w)\\ h_{0}(w)&h_{0}(w)\\ \end{array}\right)+
+(h1(w)+12r(w)00h1(w)12r(w))u+(h1(w)¯+12r(w)¯00h1(w)¯12r(w)¯)u¯++\left(\begin{array}[]{cc}h_{1}(w)+\frac{1}{2}r(w)&0\\ 0&-h_{1}(w)-\frac{1}{2}r(w)\\ \end{array}\right)u+\left(\begin{array}[]{cc}\overline{{h}_{1}(w)}+\frac{1}{2}\overline{r(w)}&0\\ 0&-\overline{{h}_{1}(w)}-\frac{1}{2}\overline{r(w)}\\ \end{array}\right)\bar{u}+
+𝒪(|u|2).+\mathcal{O}(|u|^{2}).

When acting with (Γ1,Γ1)(\Gamma_{1},-\Gamma_{1}) on both sides of the matrix (7.6), the last three terms in the final expression disappear and the result becomes, up to 𝒪(|u|2)\mathcal{O}(|u|^{2}),

(Γ1Γ1)(logλ(w)log|2u|log|2u|logλ(w))(Γ1Γ1)=\left(\begin{array}[]{cc}\Gamma_{1}&-\Gamma_{1}\par\end{array}\right)\left(\begin{array}[]{cc}\log\lambda(w)&-\log|2u|\\ -\log|2u|&\log\lambda(w)\\ \end{array}\right)\left(\begin{array}[]{cc}\Gamma_{1}\\ -\Gamma_{1}\\ \end{array}\right)= (7.7)
=2Γ12(logλ(w)+log|2u|).=2\Gamma_{1}^{2}\big{(}\log\lambda(w)+\log|2u|\big{)}.

The full Hamiltonian (7.2) therefore becomes, up to terms of order 𝒪(|u|2)\mathcal{O}(|u|^{2}),

2(w+u,wu)=Γ12π(logλ(w)+log|2u|)+2\mathcal{H}(w+u,w-u)=\frac{\Gamma_{1}^{2}}{\pi}\big{(}\log\lambda(w)+\log|2u|\big{)}+
+((a+Γ1wuw+u𝑑Uα)T,(b+Γ1wuw+u𝑑Uβ)T)(PRRTQ)(a+Γ1wuw+u𝑑Uαb+Γ1wuw+u𝑑Uβ).+\left(\begin{array}[]{cc}(a+\Gamma_{1}\int_{w-u}^{w+u}dU_{\alpha})^{T},&(b+\Gamma_{1}\int_{w-u}^{w+u}dU_{\beta})^{T}\end{array}\right)\left(\begin{array}[]{cc}P&R\\ R^{T}&Q\\ \end{array}\right)\left(\begin{array}[]{c}a+\Gamma_{1}\int_{w-u}^{w+u}dU_{\alpha}\\ b+\Gamma_{1}\int_{w-u}^{w+u}dU_{\beta}\end{array}\right).

Here one can see that the final term remains bounded as u0u\to 0, hence is negligible in this limit compared to the first term. Therefore the leading terms in this limit are given by

(w+u,wu)=Γ122π(logλ(w)+log|2u|)+𝒪(1)(u0).\mathcal{H}(w+u,w-u)=\frac{\Gamma_{1}^{2}}{2\pi}\big{(}\log\lambda(w)+\log|2u|\big{)}+\mathcal{O}(1)\quad(u\to 0). (7.8)

This is essentially a constant factor times log of the distance (in the metric) between w1=w+uw_{1}=w+u and w2=wuw_{2}=w-u. Indeed, we recover the simple and beautiful formula

(w1,w2)=Γ122πlogdist(w1,w2)+𝒪(1)(|w1w2|0)\mathcal{H}(w_{1},w_{2})=\frac{\Gamma_{1}^{2}}{2\pi}\log{\rm dist\,}(w_{1},w_{2})+\mathcal{O}(1)\quad(|w_{1}-w_{2}|\to 0)

of Boatto and Koiller. See equations (24), (25) in [3]. Compare also [19]. The distance is taken with respect to the Riemannian metric. The error term 𝒪(1)\mathcal{O}(1) can be identified with a what is called the “Batman function” in [3]. In our notations the latter is given by (10.3).

Taking the differential of (7.8) gives, on using again the metric affine connection r(w)=rmetric(w)r(w)=r_{\rm metric}(w) defined by (5.7),

d=wdw+w¯dw¯+udu+u¯du¯=d\mathcal{H}=\frac{\partial\mathcal{H}}{\partial w}dw+\frac{\partial\mathcal{H}}{\partial\bar{w}}d\bar{w}+\frac{\partial\mathcal{H}}{\partial{u}}du+\frac{\partial\mathcal{H}}{\partial\bar{u}}d\bar{u}=
=Γ124π(r(w)dw+r(w)¯dw¯+duu+du¯u¯).=\frac{\Gamma_{1}^{2}}{4\pi}\Big{(}r(w)dw+\overline{r(w)}d\bar{w}+\frac{du}{u}+\frac{d\bar{u}}{\bar{u}}\Big{)}. (7.9)

Recalling (7.5) we can expand the symplectic 22-form given by (7.1) in terms of ww and uu as

Ω=Γ12i(λ(w+u)2d(w+u)d(w¯+u¯)λ(wu)2d(wu)d(w¯u¯)=\Omega=-\frac{\Gamma_{1}}{2\mathrm{i}}\big{(}\lambda(w+u)^{2}d(w+u)\wedge d(\bar{w}+\bar{u})-\lambda(w-u)^{2}d(w-u)\wedge d(\bar{w}-\bar{u})=
=iΓ1λ(w)2(dwdu¯dw¯du+(r(w)u+r(w)u¯)(dwdw¯+dudu¯))+𝒪(|u|2).=\mathrm{i}{\Gamma_{1}}\lambda(w)^{2}\Big{(}dw\wedge d\bar{u}-d\bar{w}\wedge du+\big{(}r(w)u+\overline{r(w)u}\big{)}\big{(}dw\wedge d\bar{w}+du\wedge d\bar{u}\big{)}\Big{)}+\mathcal{O}(|u|^{2}).

With

ξ=w˙w+w¯˙w¯+u˙u+u¯˙u¯\xi=\dot{w}\frac{\partial}{\partial w}+\dot{\bar{w}}\frac{\partial}{\partial\bar{w}}+\dot{u}\frac{\partial}{\partial u}+\dot{\bar{u}}\frac{\partial}{\partial\bar{u}}

this gives, as u0u\to 0,

i(ξ)Ω=iΓ1λ(w)2((u¯˙+(r(w)u+r(w)u¯)w¯˙)dw+(u˙+(r(w)u+r(w)u¯)w˙)dw¯i(\xi)\Omega=\mathrm{i}\Gamma_{1}\lambda(w)^{2}\Big{(}-\big{(}\dot{\bar{u}}+(r(w)u+\overline{r(w)u})\dot{\bar{w}}\big{)}dw+\big{(}\dot{u}+(r(w)u+\overline{r(w)u})\dot{w}\big{)}d\bar{w}-
(w¯˙+(r(w)u+r(w)u¯)u¯˙)du+(w˙+(r(w)u+r(w)u¯)u˙)du¯+𝒪(|u|2).-\big{(}\dot{\bar{w}}+(r(w)u+\overline{r(w)u})\dot{\bar{u}}\big{)}du+\big{(}\dot{w}+(r(w)u+\overline{r(w)u}\big{)}\dot{u})d\bar{u}+\mathcal{O}(|u|^{2}\Big{)}.

Comparing with (7.9) we see that the dynamics of the vortex pair is described by the two equations

Γ14πir(w)¯=λ(w)2(u˙+(r(w)u+r(w)u¯)w˙)+𝒪(|u|2),\frac{\Gamma_{1}}{4\pi\mathrm{i}}\overline{r(w)}=\lambda(w)^{2}\big{(}\dot{{u}}+(r(w)u+\overline{r(w)u})\dot{{w}}\big{)}+\mathcal{O}(|u|^{2}), (7.10)
Γ14πi1u¯=λ(w)2(w˙+(r(w)u+r(w)u¯)u˙)+𝒪(|u|2).\frac{\Gamma_{1}}{4\pi\mathrm{i}}\frac{1}{\bar{u}}=\lambda(w)^{2}\big{(}\dot{{w}}+(r(w)u+\overline{r(w)u})\dot{{u}}\big{)}+\mathcal{O}(|u|^{2}). (7.11)

Equation (7.11) can used to eliminate w˙\dot{w} in (7.10), which then becomes

Γ14πir(w)¯=λ(w)2u˙+(r(w)u+r(w)u¯)(Γ14πiu¯λ(w)2(r(w)u+r(w)u¯)u˙)+𝒪(|u|2).\frac{\Gamma_{1}}{4\pi\mathrm{i}}\overline{r(w)}=\lambda(w)^{2}\dot{u}+\big{(}r(w)u+\overline{r(w)u}\big{)}\Big{(}\frac{\Gamma_{1}}{4\pi\mathrm{i}\bar{u}}-\lambda(w)^{2}\big{(}r(w)u+\overline{r(w)u}\big{)}\dot{u}\Big{)}+\mathcal{O}(|u|^{2}).

Here the left member cancels with one of the terms in the right member, and the rest can be written, after division by uu,

0=λ(w)2u˙u+r(w)Γ14πiu¯λ(w)2(r(w)u+r(w)u¯)2u˙u+𝒪(|u|).0=\lambda(w)^{2}\cdot\frac{\dot{u}}{u}+r(w)\frac{\Gamma_{1}}{4\pi\mathrm{i}\bar{u}}-\lambda(w)^{2}\big{(}r(w)u+\overline{r(w)u}\big{)}^{2}\cdot\frac{\dot{u}}{u}+\mathcal{O}(|u|).

In this equation the third term in the right member is of a smaller magnitude than the other two terms and can be incorporated in the final 𝒪(|u|)\mathcal{O}(|u|). Thus we arrive at

λ(w)2ddtlogu+r(w)Γ14πiu¯=0,\lambda(w)^{2}\frac{d}{dt}\log u+r(w)\frac{\Gamma_{1}}{4\pi\mathrm{i}\bar{u}}=0, (7.12)

valid with an error of at most 𝒪(|u|)\mathcal{O}(|u|) as u0u\to 0.

The above equation, (7.12), essentially comes from (7.10), and it is to be combined again with (7.11). For the latter it is enough to use the simplified form

Γ14πiu¯=λ(w)2dwdt,\frac{\Gamma_{1}}{4\pi\mathrm{i}\bar{u}}=\lambda(w)^{2}\frac{{dw}}{dt}, (7.13)

which only uses the leading terms, and for which the error is still at most 𝒪(|u|)\mathcal{O}(|u|). Inserting (7.13) in (7.12) results in the master equation

ddtlogu+r(w)dwdt=0.\frac{d}{dt}\log u+r(w)\frac{dw}{dt}=0. (7.14)

One problem with (7.14) is that the speed dw/dtdw/dt becomes infinite, along with the first term, in the dipole limit. However, this problem only affects the real part of the equation. For the imaginary part one can replace true time tt by an arbitrary parameter, which is scaled with uu so that dw/dtdw/dt remains finite as |u|0|u|\to 0. Alternatively one may scale Γ1\Gamma_{1} with uu so that the left member of (7.13) remains finite. Then one can still think of any new parameter tt as a time variable. In any case, we take imaginary parts of (7.14) and obtain

ddtargu+Im(r(w)dwdt)=0.\frac{d}{dt}\arg u+\operatorname{Im}(r(w)\frac{d{w}}{dt})=0. (7.15)

Equation (7.13) shows that the directions of uu and dw/dtdw/dt are related as

argu=argdwdt±π2,\arg u=\arg\frac{dw}{dt}\pm\frac{\pi}{2}, (7.16)

where the plus sign is to be chosen if Γ1>0\Gamma_{1}>0, the minus sign if Γ1<0\Gamma_{1}<0. Now (7.15) and (7.16) taken together give the final law for the motion of the center ww of the vortex pair in the dipole limit:

ddtargdwdt+Im(r(w)dwdt)=0.\frac{d}{dt}\arg\frac{dw}{dt}+\operatorname{Im}(r(w)\frac{d{w}}{dt})=0. (7.17)

As explained in Appendix 1, Section 10 (see in particular equation (10.5) there), (7.19) is exactly the equation for a geodesic curve when expressed in an arbitrary parameter tt. One may notice that (7.15) (and similarly for (7.17)) can be written in the parameter-free form

dargu+Im(r(w)dw)=0,{d}\arg u+\operatorname{Im}(r(w){d{w}})=0,

confirming again the fact that the geometry of the dipole trajectory has a meaning independent of any choice of parameter.

The real part of (7.14) says, in view of (10.1) (or (5.7)), that

ddt(log|u|+logλ(w))=0,\frac{d}{dt}\big{(}\log|u|+\log\lambda(w)\big{)}=0,

the error term 𝒪(|u|)\mathcal{O}(|u|) being disregarded. In other words that  |u|λ(w)=C{|u|}{\lambda(w)}=C (constant) along each trajectory. By (7.13) this also gives

|dwdt|=Cλ(w).\Big{|}\frac{dw}{dt}\Big{|}=\frac{C}{\lambda(w)}. (7.18)

Thus along each trajectory λ(w)\lambda(w) has, being proportional to one over the velocity, the same role as the refraction index in optics.

We summarize the most essential parts of the above discussion as

Theorem 7.1.

The dynamical equations for a vortex pair in the dipole limit is identical with the geodesic equation for the metric ds=λ(w)|dw|ds=\lambda(w)|dw|, namely

ddtargdwdt+Im(r(w)dwdt)=0.\frac{d}{dt}\arg\frac{dw}{dt}+\operatorname{Im}(r(w)\frac{d{w}}{dt})=0. (7.19)

Here w=w(t)w=w(t) is the location of the dipole, tt is an arbitrarily scaled time parameter chosen such that dw/dtdw/dt is finite. The orientation of the dipole is perpendicular to dw/dtdw/dt.

Remark 7.1.

It is possible to understand why dipole move along geodesics by thinking of vortex pair as a “wave front”, in an optical analogy. Equation (7.16) says that the motion is perpendicular to the wave front (the line segment from wuw-u to w+uw+u). Equation (7.15) then expresses that if the front of a vortex pair is not aligned with the level line of λ(w)\lambda(w) then the direction of uu (representing the wave front) changes in such a way that the curve w(t)w(t) bends towards higher values of λ\lambda.

Being slightly more direct and exact, on taking tt to be Euclidean arc length the first term in (7.19) is the ordinary curvature for the curve traced out by w(t)w(t). The second term can be viewed as the inner product between the gradient of λ(w)\lambda(w), which can be identified with r(w)¯\overline{r(w)}, and dw/dtdw/dt rotated 9090 degrees to the right. Letting θ\theta denote the angle between the gradient of λ(w)\lambda(w) and the velocity vector dw/dtdw/dt we can write

Im(r(w)dwdt)=|r(w)¯||dwdt|sinθ.\operatorname{Im}(r(w)\frac{d{w}}{dt})=|\overline{r(w)}||\frac{dw}{dt}|\sin{\theta}.

The above remarks are compatible with the laws of optics, for example Fermat’s law, and also with “Snell’s law” (see for example [20, 6]) in the somewhat singular case that λ(w)\lambda(w) jumps between two constant values.

8 Vortex dipoles and projective connections

A vortex dipole constructed as a vortex pair melting together is a result of two regularizations: First, a single vortex is already regularized in itself by removal of a singularity of type log|zw|\log|z-w| in the stream function, together with a possible adjustment with a term related to the curvature of the surface. This is not so severe, and everyone agrees on the result. The motion of the vortex then comes from the regular terms in the Taylor expansion, and it is easy to handle. In particular the speed of the vortex is finite.

When the two vortices in a vortex pair approach each, then a second regularization becomes necessary. If one does not want the two vortices to just annihilate each other in the limit, then one has to let the strengths of the vortices tend to infinity. Even without that, the speed of the vortex pair will be infinite in the limit, so with the vortex strengths blowing up the speed will become infinite to an even higher degree. But in any case one can speak of a residual trajectory, which will be a geodesic for the metric of the manifold, as was conjectured in [28], proved in [3], and further clarified in Theorem 7.1 above.

There is also the possibility of treating the flow directly as a dipole flow, not going via vortex pairs, and this touches on the theory of projective (or Schwarzian) connections. Below we shall sketch upon such a procedure, but it seems difficult to read off the trajectory of the dipole from this approach.

We start from the expression, using (4.25),

ν=ηdGω=(BT,AT)(dUαdUβ)ΓdG(,w)\nu=\eta-*dG^{\omega}=\left(\begin{array}[]{cc}B^{T},&-A^{T}\\ \end{array}\right)\left(\begin{array}[]{cc}dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right)-\Gamma*dG(\cdot,w)
=(bT+βdG(,w),aTαdG(,w))(dUαdUβ)ΓdG(,w)=\left(\begin{array}[]{cc}b^{T}+\oint_{\beta}*dG(\cdot,w),&-a^{T}-\oint_{\alpha}*dG(\cdot,w)\\ \end{array}\right)\left(\begin{array}[]{cc}dU_{\alpha}\\ dU_{\beta}\\ \end{array}\right)-\Gamma*dG(\cdot,w)

for the flow 11-form ν\nu in the single vortex case, and we shall to pass to the dipole limit by differentiation with respect to ww. Recall that ν\nu is a real-valued form: on writing ν=νzdz+νz¯dz¯\nu=\nu_{z}dz+\nu_{\bar{z}}d\bar{z} we have νz¯=νz¯\nu_{\bar{z}}=\overline{\nu_{z}}. When differentiating with respect to ww we get a covariant tensor, denoted dwνd_{w}\nu, which is real valued in a similar sense. In dwνd_{w}\nu there will appear two kinds of differentials: dzdz, dz¯d\bar{z} are differentials for the flow itself (as a 11-form) at a general point zz. The dipole is located at ww and the differentials dwdw, dw¯d\bar{w} represent its orientation. These two kinds of differentials are combined only via a tensor product (not an antisymmetric wedge product).

On using (3.13), (3.16), (3.17) one finds that differentiation of ν\nu with respect to ww gives, writing for simplicity dzdwdzdw in place of dzdwdz\otimes dw (etc),

dwν(z)=12j=1g(dUβj(w)dUαj(z)dUαj(w)dUβj(z))+d_{w}\nu(z)=\frac{1}{2}\sum_{j=1}^{\texttt{g}}\big{(}dU_{\beta_{j}}(w)\otimes dU_{\alpha_{j}}(z)-dU_{\alpha_{j}}(w)\otimes dU_{\beta_{j}}(z)\big{)}+
+iΓ(2G(z,w)zwdzdw+2G(z,w)zw¯dzdw¯2G(z,w)z¯wdz¯dw2G(z,w)z¯w¯dz¯dw¯).+\mathrm{i}\Gamma\big{(}\frac{\partial^{2}G(z,w)}{\partial z\partial w}dzdw+\frac{\partial^{2}G(z,w)}{\partial z\partial\bar{w}}dzd\bar{w}-\frac{\partial^{2}G(z,w)}{\partial\bar{z}\partial{w}}d\bar{z}d{w}-\frac{\partial^{2}G(z,w)}{\partial\bar{z}\partial\bar{w}}d\bar{z}d\bar{w}\big{)}.

Here the first term vanishes in the limit zwz\to w. In the remaining terms we single out the poles as separate terms and let zwz\to w in what remains. This gives, on using (11.6), (11.7), (11.5) and representing the first term simply by 𝒪(|zw|)\mathcal{O}(|z-w|) (at the end),

dwν(z)=Γ4πi((1(zw)2+22H(z,w)zw)dzdw+22H(z,w)zw¯dzdw¯d_{w}\nu(z)=-\frac{\Gamma}{4\pi\mathrm{i}}\Big{(}\big{(}\frac{1}{(z-w)^{2}}+2\frac{\partial^{2}H(z,w)}{\partial z\partial w}\big{)}dzdw+2\frac{\partial^{2}H(z,w)}{\partial z\partial\bar{w}}dzd\bar{w}-
22H(z,w)z¯wdz¯dw(1(z¯w¯)2+22H(z,w)z¯w¯)dz¯dw¯)+𝒪(|zw|)=-2\frac{\partial^{2}H(z,w)}{\partial\bar{z}\partial{w}}d\bar{z}d{w}-\big{(}\frac{1}{(\bar{z}-\bar{w})^{2}}+2\frac{\partial^{2}H(z,w)}{\partial\bar{z}\partial\bar{w}}\big{)}d\bar{z}d\bar{w}\Big{)}+\mathcal{O}(|z-w|)=
=Γ2πIm((1(zw)2+h1(w)w2h2(w))dzdw(h1(w)w¯2h11(w)))dzdw¯)=-\frac{\Gamma}{2\pi}\operatorname{Im}\Big{(}\big{(}\frac{1}{(z-w)^{2}}+\frac{\partial h_{1}(w)}{\partial w}-2h_{2}(w)\big{)}dzdw-\big{(}\frac{\partial h_{1}(w)}{\partial\bar{w}}-2h_{11}(w)\big{)})dzd\bar{w}\Big{)}
+𝒪(|zw|).+\mathcal{O}(|z-w|).

To get the speed of dipole one would like to insert z=wz=w above, but clearly this does not give any sensible result, just a flow 11-form in zz that becomes infinite at z=wz=w. Still this singular flow is associated with a certain direction determined by dwdw, or more precisely by the action of dwdw on a vector m representing the orientation of the dipole. See Example 8.1 below. In any case, this singular term overrules all other terms.

Up to a real factor, we thus have above the singularity

Imdzdw(zw)2\operatorname{Im}\frac{dzdw}{(z-w)^{2}}

plus the finite quantity

Im((h1(w)w2h2(w))dzdw(h1(w)w¯2h11(w))dzdw¯).\operatorname{Im}\Big{(}\big{(}\frac{\partial h_{1}(w)}{\partial w}-2h_{2}(w)\big{)}dzdw-\big{(}\frac{\partial h_{1}(w)}{\partial\bar{w}}-2h_{11}(w)\big{)}dzd\bar{w}\Big{)}. (8.1)

Certainly this finite term will be of minor importance compared with the singular term. The singular term actually determines the entire flow (at a given instance) once the orientation of the dipole is given and, for example, the imaginary parts of the periods are given, say are set to zero. Indeed, on letting dwdw act on a vector m representing the orientation, dw,m\langle dw,{\texttt{m}}\rangle becomes a complex number and there remains the differentials dzdz and dz¯d\bar{z} representing the flow, and this can then be associated with with a unique meromorphic differential on MM.

To elaborate the above statements a little, let m\texttt{m}\in{\mathbb{C}} be given. Then there exists a unique meromorphic differential μ=μzdz\mu=\mu_{z}dz on MM of the form

μ=mdz(zw)2+regular\mu=\frac{{\texttt{m}}\,dz}{(z-w)^{2}}+{\rm regular}

and having periods

Imαkμ=0,Imβkμ=0(k=1,,g).\operatorname{Im}\oint_{\alpha_{k}}\mu=0,\quad\operatorname{Im}\oint_{\beta_{k}}\mu=0\quad(k=1,\dots,\texttt{g}).

Such a μ\mu represents the dipole flow via

dwν,m=Imμ,\langle d_{w}\nu,\cdot\otimes{\texttt{m}}\rangle=\operatorname{Im}\mu,

where the left member means that the differentials dwdw and dw¯d\bar{w} in dwνd_{w}\nu shall act on m. In practice this just means that dwdw and dw¯d\bar{w} shall be replaced by m and m¯\bar{\texttt{m}}, respectively, regarded as complex numbers

However, this purely conformal picture (we have not yet used the metric) gives no information of how the dipole is to move. That must be determined by the metric, and there seems to be no other reasonable possibility than that it shall move along the geodesic perpendicular to m. In particular, the regular part (8.1) seems not to influence the dipole dynamics. What one can say from a mathematical point of view is that the coefficient in the leading term of (8.1),

16qrobin(w)=h1(w)w2h2(w)-\frac{1}{6}q_{\rm robin}(w)=\frac{\partial h_{1}(w)}{\partial w}-2h_{2}(w)

is a projective connection, up to a factor, see Lemma 11.2. The second term in (8.1) is directly linked to the two metrics involved: by (6.2), (6.1), (11.5) we have

h1(w)w¯2h11(w)=14Δh0(w)πVλ(w)2.\frac{\partial h_{1}(w)}{\partial\bar{w}}-2h_{11}(w)=\frac{1}{4}\Delta h_{0}(w)-\frac{\pi}{V}\lambda(w)^{2}.

This entire quantity transforms as the density of a metric. But it is not necessarily positive. In the case of a sphere, for example, it vanishes (see Example 6.1). Also qrobinq_{\rm robin} vanishes in the case.

Example 8.1.

As an example, let the initial condition be that w=0w=0 and m=v{\texttt{m}}=\frac{\partial}{\partial v}, where w=u+ivw=u+\mathrm{i}v. Then dw,m=i\langle dw,{\texttt{m}}\rangle=\mathrm{i}, and the flow 11-form from the singularity becomes

Imdzdw,m(zw)2=Redzz2=(x2y2)dx2xydy(x2+y2)2.\operatorname{Im}\frac{dz\,\langle dw,{\texttt{m}}\rangle}{(z-w)^{2}}=\operatorname{Re}\frac{dz}{z^{2}}=\frac{(x^{2}-y^{2})dx-2xydy}{(x^{2}+y^{2})^{2}}.

On the xx-axis (y=0y=0) this is dx/x2{dx}/{x^{2}}, in other words a flow along the axis with a polar singularity of order two, as expected. Indeed, the entire flow is the well-known dipole picture which appears in many applications of conformal mapping.

9 Remarks on vortex motion in planar domains

Vortex motion in a planar domain can easily be treated as a special case of vortex motion on Riemann surfaces by turning to the Schottky double of the planar domain. For simplicity we shall only discuss the case of one single vortex in the planar domain. The case of several vortices will be similar in an obvious way. The ideas in this section extend to cases of vortex motion on general open Riemannian surfaces with analytic boundary.

Let Ω\Omega\subset{\mathbb{C}} be the planar domain, assumed to be bounded by finitely many real analytic curves. The Schottky double, first described in [45], is the compact Riemann surface M=Ω^M=\hat{\Omega} obtained by completing Ω\Omega with a “backside” Ω~\tilde{\Omega}, having the opposite conformal structure, and glueing the two along the common boundary. Thus Ω^=ΩΩΩ~\hat{\Omega}=\Omega\cup\partial\Omega\cup\tilde{\Omega} in a set theoretic sense, and the conformal structure becomes smooth over Ω\partial\Omega, as can be seen from well-known reflection principles. If zz is a point in ΩM\Omega\subset M, then z~\tilde{z} will denote the corresponding (reflected) point on Ω~M\tilde{\Omega}\subset M. Both zz and z~\tilde{z} can also be considered as points in {\mathbb{C}}, then serving as coordinates of the mentioned points in MM (holomorphic respectively anti-holomorphic coordinates), and as such they are the same: z=z~z=\tilde{z}\in{\mathbb{C}}.

In our case we need also a Riemannian structure, with a metric. This is to be the Euclidean metric on each of Ω\Omega and Ω~\tilde{\Omega}, but these do not fit smoothly across curved parts of Ω\partial\Omega, it will only be Lipschitz continuous. But that is good enough for our purposes because the vortex will anyway never approach the boundary. (In the case of several vortices it is however possible to make up situations in which vortices do reach the boundary).

The metric on M=Ω^M=\hat{\Omega} is thus to be

ds={|dz|,zΩ,|dz~|,z~Ω~.ds=\begin{cases}|dz|,\quad z\in\Omega,\\ |d\tilde{z}|,\quad\tilde{z}\in\tilde{\Omega}.\end{cases} (9.1)

To see how this behaves across Ω\partial\Omega we need a holomorphic coordinate defined in a full neighborhood of this curve in MM. A natural candidate can be defined in terms of the Schwarz function S(z)S(z) for Ω\partial\Omega, a function which is defined by its properties of being holomorphic in a neighborhood of Ω\partial\Omega in {\mathbb{C}} and by satisfying

S(z)=z¯on Ω.S(z)=\bar{z}\quad\text{on }\partial\Omega. (9.2)

See [9, 47] for details about S(z)S(z). We remark that zS(z)¯z\mapsto\overline{S(z)} is the (local) anti-conformal reflection map in Ω\partial\Omega and that S(z)=T(z)2S^{\prime}(z)=T(z)^{-2}, where T(z)T(z) is the positively oriented and holomorphically extended unit tangent vector on Ω\partial\Omega.

The complex coordinate zz in Ω\Omega extends, as a holomorphic function, to a full neighborhood of ΩΩ\Omega\cup\partial\Omega, both when this neighborhood is considered as a subset of {\mathbb{C}} and when it is considered as a subset of MM. The first case is trivial, and the second case depends on Ω\partial\Omega being analytic. In terms of the Schwarz function this second extension is given by

z={zfor zΩΩ,S(z~)¯for z~Ω~, close to Ω.z=\begin{cases}z\quad&\text{for }z\in\Omega\cup\partial\Omega,\\ \overline{S({\tilde{z}})}\quad&\text{for }\tilde{z}\in\tilde{\Omega},\text{ close to }\partial\Omega.\end{cases} (9.3)

In the latter expression, S(z~)¯\overline{S(\tilde{z})}, z~\tilde{z} is to be interpreted as a complex number. When the metric on MM is expressed in the coordinate (9.3) it becomes

ds={|dz|for zΩΩ,|S(z)||dz|for zΩ¯, close to Ω.ds=\begin{cases}|dz|\quad&\text{for }z\in\Omega\cup\partial\Omega,\\ |S^{\prime}({z})||d{z}|\quad&\text{for }{z}\in{\mathbb{C}}\setminus\overline{\Omega},\text{ close to }\partial\Omega.\end{cases} (9.4)

In the second case, z=S(z~)¯z=\overline{S(\tilde{z})}, z~Ω~\tilde{z}\in\tilde{\Omega}, whereby z~=S(z)¯\tilde{z}=\overline{S(z)} and so |dz~|=|S(z)||dz||d\tilde{z}|=|S^{\prime}({z})||d{z}|. Thus (9.4) is consistent with (9.1). We see from the coordinate representation (9.4) that the metric is only Lipschitz continuous across Ω\partial\Omega. This is the best one can expect.

The associated affine connection (5.7) (or (10.1)) is in the coordinate (9.3) given by

r(z)={0for zΩΩ,{S(z),z}1for zΩ¯, close to Ω,r(z)=\begin{cases}0\quad&\text{for }z\in\Omega\cup\partial\Omega,\\ \{S({z}),{z}\}_{1}\quad&\text{for }{z}\in{\mathbb{C}}\setminus\overline{\Omega},\text{ close to }\partial\Omega,\end{cases}

where (see Appendix 1, Section 10, for notations)

{S(z),z}1=S′′(z)S(z)=2T(z)T(z).\{S(z),z\}_{1}=\frac{S^{\prime\prime}(z)}{S^{\prime}(z)}=-2\frac{T^{\prime}(z)}{T(z)}. (9.5)

Thus r(z)r(z) is discontinuous across Ω\partial\Omega and it should on this curve be represented by its mean-value

rMV(z)=12{S(z),z}1=T(z)T(z)(zΩ).r_{\rm MV}(z)=\frac{1}{2}\{S(z),z\}_{1}=-\frac{T^{\prime}(z)}{T(z)}\quad(z\in\partial\Omega). (9.6)
Example 9.1.

Let Ω=𝔻\Omega={\mathbb{D}}. Then S(z)=1zS(z)=\frac{1}{z} and the coordinate zz in (9.3) extends to the entire complex plane, thus representing all of M=𝔻𝔻𝔻~M={\mathbb{D}}\cup\partial{\mathbb{D}}\cup\tilde{{\mathbb{D}}} except for the point 0~𝔻~\tilde{0}\in\tilde{{\mathbb{D}}}. And the metric expressed in this coordinate becomes

ds={|dz|,|z|1,|z|2|dz|,|z|>1.ds=\begin{cases}|dz|,\quad&|z|\leq 1,\\ |z|^{-2}{|d{z}|},\quad&|z|>1.\end{cases}

The affine connection similarly becomes, including the mean-value on the boundary,

r(z)={0,|z|<1,z,|z|=1,2z,|z|>1.r(z)=\begin{cases}0,\quad&|z|<1,\\ -z,\quad&|z|=1,\\ -2z,\quad&|z|>1.\end{cases}

The geodesic curves in Ω\Omega are of course the (Euclidean) straight lines (similarly in Ω~\tilde{\Omega}), geodesic curves crossing Ω\partial\Omega are straight lines reflecting into the other side under equal angles on Ω\partial\Omega (just as ordinary optical reflection), while Ω\partial\Omega is in itself a geodesic curve. The latter is intuitively obvious since at any point on Ω\partial\Omega there should be one geodesic in the tangential direction, and this has no other way to go than to follow the boundary.

To confirm the last statement analytically, let tt be an arc length (with respect to the Euclidean metric) parameter along Ω\partial\Omega, so that T(z)=dzdtT(z)=\frac{dz}{dt} on Ω\partial\Omega. The curvature κ\kappa of Ω\partial\Omega is

κ=ddtargdzdtzΩ,\kappa=\frac{d}{dt}\arg\frac{dz}{dt}\quad z\in\partial\Omega,

and using that T(z)T(z)¯=1T(z)\overline{T(z)}=1 on Ω\partial\Omega one finds that

T(z)=iκ(zΩ).T^{\prime}(z)=\mathrm{i}\kappa\quad(z\in\partial\Omega).

In particular T(z)T^{\prime}(z), and so rMV(z)T(z)r_{\rm MV}(z)T(z) (by (9.6)), is purely imaginary on Ω\partial\Omega. Combining with (9.6) it follows that

ddtargdzdt=irMV(z)T(z),\frac{d}{dt}\arg\frac{dz}{dt}={\mathrm{i}}r_{\rm MV}(z)T(z),

hence

ddtargdzdt+Im(rMV(z)T(z))=0(zΩ).\frac{d}{dt}\arg\frac{dz}{dt}+{\operatorname{Im}}\big{(}r_{\rm MV}(z)T(z)\big{)}=0\quad(z\in\partial\Omega).

Thus the geodesic equation (7.19) holds for the curve Ω\partial\Omega, as claimed.

We remark also that the curvature κ\kappa of the boundary curve Ω\partial\Omega appears also in the expression for the Gaussian curvature for the metric on MM. That curvature vanishes on Ω\Omega and on Ω~\tilde{\Omega}, whereas it on Ω\partial\Omega has a singular contribution, with density 2κ2\kappa with respect arc-length measure on Ω\partial\Omega.

A single vortex in a planar domain Ω\Omega moves along a level line of the appropriate Robin function, or Routh’s stream function [34, 35, 36]). If the vortex is close to the boundary then it follows the boundary closely, with high speed. From the perspective of the Schottky double the boundary conditions for planar fluid motion are such that there is automatically a mirror vortex on the other side in the double, thus we really have a vortex pair close to Ω\partial\Omega on the double. In the limit this becomes a vortex dipole, moving with infinite speed along the geodesic Ω\partial\Omega.

Considering in some more detail such a symmetric vortex pair, with vortex locations wΩw\in\Omega and w~Ω~\tilde{w}\in\tilde{\Omega}, we first notice that the Green function Gω(z)G^{\omega}(z) for ω=δwδw~\omega=\delta_{w}-\delta_{\tilde{w}} simply is the anti-symmetric extension to the Schottky double of the ordinary Green function GΩ(z,w)G_{\Omega}(z,w) for Ω\Omega:

Gδwδw~(z)=GΩ(z,w)(zΩ).G^{\delta_{w}-\delta_{\tilde{w}}}(z)=G_{\Omega}(z,w)\quad(z\in\Omega).

Then the stream function ψ\psi in (4.2) becomes what is sometimes called the hydrodynamic Green function, which depends on the prescribed periods. This function, which can be traced back (at least in special cases) to [29, 36], has more recently been discussed in for example [8, 14, 13, 25].

We wish to clarify how this hydrodynamic Green function is related to the modification, in the beginning of Section 4, of the general Green function flow dGω-*dG^{\omega} by an additional term η\eta. To do this we fix, in the case of a Schottky double M=Ω^M=\hat{\Omega}, the homology basis in such a way that the curves βj\beta_{j}, j=1,,gj=1,\dots,\texttt{g}, closely follow the inner components of Ω\partial\Omega, and each curve αj\alpha_{j} goes from the jj:th inner component of Ω\partial\Omega through Ω\Omega to the outer component, and then back again on the backside.

We also introduce the harmonic measures uju_{j}, j=1,,gj=1,\dots,\texttt{g}, here defined to be those harmonic functions in Ω\Omega which takes the boundary value one on the designated (number jj) inner component of Ω\partial\Omega and vanishes on the rest of Ω\partial\Omega. Their differentials dujdu_{j} extend harmonically to the Schottky double with αk𝑑uj=2δkj\oint_{\alpha_{k}}du_{j}=-2\delta_{kj}, αk𝑑uj=0\oint_{\alpha_{k}}du_{j}=0. Thus duj=2dUβjdu_{j}=-2dU_{\beta_{j}} in terms of our general notations (as in (3.18), (3.19)).

In the block matrix notation of (4.19) we have

dUβ=PdUα+RdUβ,-*dU_{\beta}=P\,dU_{\alpha}+R\,dU_{\beta},

where P=(Pkj)P=(P_{kj}), R=(Rkj)R=(R_{kj}) and (see (3.18), (3.19))

Pkj=βjdUβk,Rkj=αjdUβk.P_{kj}=-\oint_{\beta_{j}}*dU_{\beta_{k}},\quad R_{kj}=\oint_{\alpha_{j}}*dU_{\beta_{k}}.

The last integral can be written

Rkj=12αjduk=12αjΩukn𝑑s12αjΩ~ukn𝑑s,R_{kj}=-\frac{1}{2}\oint_{\alpha_{j}}*du_{k}=-\frac{1}{2}\int_{\alpha_{j}\cap\Omega}\frac{\partial u_{k}}{\partial n}ds-\frac{1}{2}\int_{\alpha_{j}\cap\tilde{\Omega}}\frac{\partial u_{k}}{\partial n}ds,

and it is easy to see that it is zero because of cancelling contributions from Ω\Omega and Ω~\tilde{\Omega} due to the symmetry of dukdu_{k} and αj\alpha_{j} going the opposite way on the backside.

As a general conclusion we therefore have that R=0R=0 in the matrix (4.19) when MM is the Schottky double of a planar domain. As a consequence,

dUβ=PdUα.-*dU_{\beta}=P\,dU_{\alpha}.

Similarly, the other equation contained in (4.19) becomes

dUα=QdUβ.*dU_{\alpha}=Q\,dU_{\beta}. (9.7)

Turning now to flow η\eta in (4.15), this is necessarily symmetric on M=Ω^M=\hat{\Omega}, hence

αjη=0.\oint_{\alpha_{j}}\eta=0.

It follows that the coefficients AjA_{j} in (4.15) vanish, whereby that equation becomes

η=j=1gBjdUαj.\eta=\sum_{j=1}^{\texttt{g}}B_{j}dU_{\alpha_{j}}.

In terms of the stream function ψ=Gω+ψ0\psi=G^{\omega}+\psi_{0} (see again (4.2)) this gives, inserting also (9.7),

dψ0=η=j=1gBjdUαj=j=1gCjdUβj=d(12j=1gCjuj),d\psi_{0}=*\eta=\sum_{j=1}^{\texttt{g}}B_{j}*dU_{\alpha_{j}}=\sum_{j=1}^{\texttt{g}}C_{j}dU_{\beta_{j}}=d\big{(}-\frac{1}{2}\sum_{j=1}^{\texttt{g}}C_{j}u_{j}\big{)},

with Cj=i=1gBiQijC_{j}=\sum_{i=1}^{\texttt{g}}B_{i}Q_{ij}. It follows in particular that ψ0\psi_{0}, and hence all of ψ\psi, is single-valued on Ω\Omega.

In summary, the total stream function is

ψ(z)=GΩ(z,w)+j=1gCjUβj(z),\psi(z)=G_{\Omega}({z,w})+\sum_{j=1}^{\texttt{g}}C_{j}U_{\beta_{j}}(z),

and it is single-valued when restricted to Ω\Omega. It is clear that the CjC_{j} will actually depend on ww, and for symmetry reasons the above formula eventually takes the form

ψ(z)=GΩ(z,w)+i,j=1gCijUβi(z)Uβj(w)\psi(z)=G_{\Omega}({z,w})+\sum_{i,j=1}^{\texttt{g}}C_{ij}U_{\beta_{i}}(z)U_{\beta_{j}}(w)

for some constants CijC_{ij} subject to Cij=CjiC_{ij}={C_{ji}}.

10 Appendix 1: Affine and projective connections

Besides differential forms, and tensor fields in general, affine and projective connections are quantities on Riemann surfaces which are relevant for point vortex motion. Therefore we give below a short introduction to these notions. The affine connections have the same meanings as in ordinary differential geometry, used to define covariant derivatives for example, and they play an important role in many areas of mathematical physics. Projective connections have more recently become important, in for example conformal field theory.

Some general references for the kind of connections we are going to consider are [43, 21, 22, 23, 25]. We define them in the simplest possible manner, namely as quantities defined in terms of local holomorphic coordinates and transforming in specified ways when changing from one coordinate to another.

Let z~=φ(z)\tilde{z}=\varphi(z) represent a holomorphic local change of complex coordinate on a Riemann surface MM and define three nonlinear differential expressions {,}k\{\cdot,\cdot\}_{k}, k=0,1,2k=0,1,2, by

{z~,z}0\displaystyle\{\tilde{z},z\}_{0} =logφ(z)\displaystyle=\log\varphi^{\prime}(z) =2log1φ\displaystyle=-2\log\frac{1}{\sqrt{\varphi^{\prime}}}
{z~,z}1\displaystyle\{\tilde{z},z\}_{1} =(logφ(z))=φ′′φ\displaystyle=(\log\varphi^{\prime}(z))^{\prime}=\frac{\varphi^{\prime\prime}}{\varphi^{\prime}} =2φ(1φ)\displaystyle=-2\sqrt{\varphi^{\prime}}\,\,(\frac{1}{\sqrt{\varphi^{\prime}}})^{\prime}
{z~,z}2\displaystyle\{\tilde{z},z\}_{2} =(logφ(z))′′12((logφ(z)))2=φ′′′φ32(φ′′φ)2\displaystyle=(\log\varphi^{\prime}(z))^{\prime\prime}-\frac{1}{2}((\log\varphi^{\prime}(z))^{\prime})^{2}=\frac{\varphi^{\prime\prime\prime}}{\varphi^{\prime}}-\frac{3}{2}(\frac{\varphi^{\prime\prime}}{\varphi^{\prime}})^{2} =2φ(1φ)′′\displaystyle=-2\sqrt{\varphi^{\prime}}\,\,(\frac{1}{\sqrt{\varphi^{\prime}}})^{\prime\prime}

The last expression is the Schwarzian derivative of φ\varphi. For {z~,z}0\{\tilde{z},z\}_{0} there is an additive indetermincy of 2πi2\pi\mathrm{i}, so actually only its real part, or exponential, is completely well-defined.

As alternative definitions we have, with z~=φ(z)\tilde{z}=\varphi(z), w~=φ(w)\tilde{w}=\varphi(w),

{w~,w}0\displaystyle\{\tilde{w},w\}_{0} =limzwlogz~w~zw,\displaystyle=\lim_{z\to w}\log\frac{\tilde{z}-\tilde{w}}{z-w},
{w~,w}1\displaystyle\{\tilde{w},w\}_{1} =2limzwzlogz~w~zw,\displaystyle=2\lim_{z\to w}\frac{\partial}{\partial z}\log\frac{\tilde{z}-\tilde{w}}{z-w},
{w~,w}2\displaystyle\{\tilde{w},w\}_{2} =6limzw2zwlogz~w~zw.\displaystyle=6\lim_{z\to w}\frac{\partial^{2}}{\partial z\partial w}\log\frac{\tilde{z}-\tilde{w}}{z-w}.

The following chain rules hold, if zz depends on ww via an intermediate variable uu:

{z,w}k(dw)k={z,u}k(du)k+{u,w}k(dw)k(k=0,1,2).\{z,w\}_{k}(dw)^{k}=\{z,u\}_{k}(du)^{k}+\{u,w\}_{k}(dw)^{k}\quad(k=0,1,2).

In particular,

{z,w}k(dw)k={w,z}k(dz)k(k=0,1,2).\{z,w\}_{k}(dw)^{k}=-\{w,z\}_{k}(dz)^{k}\quad(k=0,1,2).

It turns out that the three operators {,}k\{\cdot,\cdot\}_{k}, k=0,1,2k=0,1,2, are unique in having properties as above, i.e., one cannot go on with anything similar for k3k\geq 3. See [21, 22] for details.

Definition 10.1.

An affine connection (or 11-connection) on MM is an object which is represented by local differentials r(z)dzr(z)dz, r~(z~)dz~\tilde{r}(\tilde{z})d\tilde{z},…(one in each coordinate variable, and not necessarily holomorphic) glued together according to the rule

r~(z~)dz~=r(z)dz{z~,z}1dz.\tilde{r}({\tilde{z}}){d\tilde{z}}=r(z){dz}-\{\tilde{z},z\}_{1}\,{dz}.
Definition 10.2.

A projective connection (or Schwarzian connection, or 22-connection) on MM consists of local quadratic differentials q(z)(dz)2q(z)(dz)^{2}, q~(z~)(dz~)2\tilde{q}(\tilde{z})(d\tilde{z})^{2}, …, glued together according to

q~(z~)(dz~)2=q(z)(dz)2{z~,z}2(dz)2.\tilde{q}({\tilde{z}})({d\tilde{z}})^{2}=q(z)({dz})^{2}-\{\tilde{z},z\}_{2}\,({dz})^{2}.

One may also consider 0-connections, assumed here to be real-valued. Such a connection p(z)p(z) transforms according to

p~(z~)=p(z)Re{z~,z}0.\tilde{p}(\tilde{z})=p(z)-\operatorname{Re}\{\tilde{z},z\}_{0}.

This means exactly that

ds=ep(z)|dz|.ds=e^{p(z)}{|dz|}.

is a Riemannian metric.

For a metric in general, ds=λ(z)|dz|=ep(z)|dz|ds=\lambda(z){|dz|}=e^{p(z)}|dz|, there is a natural affine connection associated to it by

r(z)=2zlogλ(z)=2pz=pxipy.r(z)=2\frac{\partial}{\partial z}\log\lambda(z)=2\frac{\partial p}{\partial z}=\frac{\partial p}{\partial x}-\mathrm{i}\frac{\partial p}{\partial y}. (10.1)

This can be identified with the Levi-Civita connection in general tensor analysis. The real and imaginary parts coincide (up to sign) with the components of the classical Christoffel symbols Γijk\Gamma_{ij}^{k}. The Gaussian curvature of the metric is

κ(z)=4λ(z)22zz¯logλ(z)=2λ(z)2r(z)z¯.\kappa(z)=-4\lambda(z)^{-2}\frac{\partial^{2}}{\partial z\partial\bar{z}}\log\lambda(z)=-2{\lambda(z)^{-2}}\frac{\partial r(z)}{\partial\bar{z}}.

Under conformal changes of coordinates κ\kappa transforms as a scalar: κ~(z~)=κ(z)\tilde{\kappa}(\tilde{z})=\kappa(z) in previous notation.

Independent of any metric, an affine connection rr gives rise to a projective connection qq by

q(z)=r(z)z12r(z)2.q(z)=\frac{\partial r(z)}{\partial z}-\frac{1}{2}r(z)^{2}. (10.2)

This qq is sometimes called the “curvature” of rr (see [11]). That curvature is however not the same as the Gaussian curvature in case r(z)r(z) comes form a metric. A polarized version of qq is

q(z,w)=12(r(z)z+r(w)wr(z)r(w)),q(z,w)=\frac{1}{2}\big{(}\frac{\partial r(z)}{\partial z}+\frac{\partial r(w)}{\partial w}-r(z)r(w)\big{)}, (10.3)

related to the Batman function in [3].

The equation for geodesic curves z=z(t)z=z(t) is, in terms of an arc-length parameter tt,

d2zdt2+r(z)(dzdt)2=0\frac{d^{2}z}{dt^{2}}+r(z)(\frac{dz}{dt})^{2}=0

or, written in another way,

ddtlogdzdt+r(z)dzdt=0.\frac{d}{dt}\log\frac{dz}{dt}+r(z)\frac{dz}{dt}=0. (10.4)

The first version is just a reformulation of the usual equation in terms of Christoffel functions in ordinary differential geometry (see [15]).

The real part of (10.4) only contains information about the parametrization, while the imaginary part, namely

ddtargdzdt+Im(r(z)dzdt)=0.\frac{d}{dt}\arg\frac{dz}{dt}+\operatorname{Im}\big{(}r(z)\frac{dz}{dt}\big{)}=0. (10.5)

describes the geodesic curve in terms of an arbitrary real parameter tt. The latter property is useful in the context of the motion of vortex pairs since the speed of such a pair tends to infinity as the distance between the two vortices goes zero, and time therefore has to be successively reparametrized.

11 Appendix 2: Behavior of singular parts under changes of coordinates

The stream function ψ\psi for a flow in a neighborhood of a point vortex with location ww has an expansion starting, in local coordinates and up to a constant factor depending on the strength of the vortex,

ψ(z)=log|zw|c0(w)𝒪(|zw|).\psi(z)=\log|z-w|-c_{0}(w)-\mathcal{O}(|z-w|).

Similarly, the expansion of the analytic completion ν+iν=2iψzdz\nu+\mathrm{i}*\nu=2\mathrm{i}\frac{\partial\psi}{\partial{z}}dz of the flow 11-form ν\nu is (again up to a constant factor)

f(z)dz=dzzwc1(w)dz𝒪(zw).f(z)dz=\frac{dz}{z-w}-c_{1}(w)dz-\mathcal{O}(z-w).

One step further, one may consider a pure vortex dipole. The corresponding flow is obtained by differentiating the above ν=Re(f(z)dz)\nu=\operatorname{Re}\big{(}f(z)dz\big{)} with respect to aa. The analytic completion F(z)dzdwF(z)dzdw of the so obtained flow 11-form dwνd_{w}\nu has an expansion

F(z)dzdw=dzdw(zw)22c2(w)dzdw+𝒪(zw).F(z)dzdw=\frac{dzdw}{(z-w)^{2}}-2c_{2}(w)dzdw+\mathcal{O}(z-w).

Above, and always in similar situations, dzdwdzdw should be interpreted as the tensor product dzdwdz\otimes dw, i.e. it is not a wedge product.

The coefficients c0c_{0}, c1c_{1}, c2c_{2} above transforms, under conformal changes of coordinates, as different kinds of connections as defined in Appendix, Section 10. This is made precise in the following elementary lemma, which we here cite without proof from [24]. Compare Lemma 3.6 and subsequent discussions in [2].

Lemma 11.1.

Let z~=φ(z)\tilde{z}=\varphi(z) be a local conformal map representing a change of coordinates near a point z=wz=w, and set w~=φ(w)\tilde{w}=\varphi(w). Let ψ(z)\psi(z) (=ψ(z,w)=\psi(z,w)) be a locally defined real-valued harmonic function with a logarithmic pole at z=wz=w, similarly f(z)dzf(z)dz a meromorphic differential with a simple pole at the same point, and F(z)dzdwF(z)dzdw a double differential with a pure (residue free) second order pole. Precisely, we assume the following local forms, in the zz and z~\tilde{z} variables:

ψ(z)\displaystyle\psi(z) =log|zw|c0(w)+𝒪(|zw|)\displaystyle=\log|z-w|-c_{0}(w)+\mathcal{O}(|{z}-{w}|)
=log|z~w~|c~0(w~)+𝒪(|z~w~|),\displaystyle=\log|\tilde{z}-\tilde{w}|-\tilde{c}_{0}(\tilde{w})+\mathcal{O}(|\tilde{z}-\tilde{w}|),
f(z)dz\displaystyle f(z)dz =dzzwc1(w)dz+𝒪(zw)\displaystyle=\frac{dz}{z-w}-c_{1}(w)dz+\mathcal{O}(z-w)
=dz~z~w~c~1(w~)dz~+𝒪(z~w~),\displaystyle=\frac{d\tilde{z}}{\tilde{z}-\tilde{w}}-\tilde{c}_{1}(\tilde{w})d\tilde{z}+\mathcal{O}(\tilde{z}-\tilde{w}),
F(z)dzdw\displaystyle F(z)dzdw =dzdw(zw)22c2(w)dzdw+𝒪(zw)\displaystyle=\frac{dzdw}{(z-w)^{2}}-2c_{2}(w)dzdw+\mathcal{O}(z-w)
=dz~dw~(z~w~)22c~2(w~)dz~dw~+𝒪(z~w~).\displaystyle=\frac{d\tilde{z}d\tilde{w}}{(\tilde{z}-\tilde{w})^{2}}-2\tilde{c}_{2}(\tilde{w})d\tilde{z}d\tilde{w}+\mathcal{O}(\tilde{z}-\tilde{w}).

Then, as functions of the location of the singularity and up to constant factors, c0c_{0} transforms as a 0-connection, c1c_{1} as an affine connection and c2c_{2} as a projective connection:

c~0(w~)=c0(w)+Re{w~,w}0,\tilde{c}_{0}(\tilde{w})=c_{0}(w)+\operatorname{Re}\{\tilde{w},w\}_{0},
c~1(w~)dw~=c1(w)dw+12{w~,w}1dw,\tilde{c}_{1}(\tilde{w})d\tilde{w}=c_{1}(w)dw+\frac{1}{2}\{\tilde{w},w\}_{1}dw,
2c~2(w~)dw~2=2c2(w)dw2+16{w~,w}2dw2.2\tilde{c}_{2}(\tilde{w})d\tilde{w}^{2}=2c_{2}(w)dw^{2}+\frac{1}{6}\{\tilde{w},w\}_{2}dw^{2}.

The first statement is most conveniently expressed by saying that

ds=ec~0(w~)|dw~|=ec0(w)|dw|ds=e^{-\tilde{c}_{0}(\tilde{w})}|d\tilde{w}|=e^{-c_{0}(w)}|dw|

defines a conformally invariant metric.

Remark 11.1.

It is actually not necessary for the conclusions of the lemma that ψ\psi, fdzfdz, FdzdwFdzdw are harmonic/analytic away from the singularity, it is enough that the local forms of the singularity and constant terms given as above hold.

We may adapt the above lemma to the Green function G(z,w)=Gδw(z)G(z,w)=G^{\delta_{w}}(z), despite it is not harmonic in zz (because of the compensating background flow). First we write

G(z,w)=12π(log|zw|+H(z,w))G(z,w)=\frac{1}{2\pi}(-\log|z-w|+H(z,w)) (11.1)

and then expand the regular part as

H(z,w)=h0(w)+12(h1(w)(zw)+h1(w)¯(z¯w¯))+H(z,w)=h_{0}(w)+\frac{1}{2}\left(h_{1}(w)(z-w)+\overline{h_{1}(w)}(\bar{z}-\bar{w})\right)+ (11.2)
+12(h2(w)(zw)2+h2(w)¯(z¯w¯)2)+h11(w)(zw)(z¯w¯)+𝒪(|zw|3).+\frac{1}{2}\left(h_{2}(w)(z-w)^{2}+\overline{h_{2}(w)}(\bar{z}-\bar{w})^{2}\right)+h_{11}(w)(z-w)(\bar{z}-\bar{w})+\mathcal{O}(|z-w|^{3}).

We note from (4.10) that

H(w,w)=h0(w),{H(z,w)z}z=w=12h1(w).H(w,w)=h_{0}(w),\quad\{\frac{\partial H(z,w)}{\partial z}\}_{z=w}=\frac{1}{2}h_{1}(w).

Thus the symmetry of H(z,w)H(z,w) gives

h1(w)=h0(w)w,h_{1}(w)=\frac{\partial h_{0}(w)}{\partial w}, (11.3)

For the second order derivatives we have

{2H(z,w)z2}z=w=h2(w),\{\frac{\partial^{2}H(z,w)}{\partial z^{2}}\}_{z=w}=h_{2}(w),\qquad\qquad\quad (11.4)
{2H(z,w)zz¯}z=w=h11(w)=π2Vλ(w)2,\{\frac{\partial^{2}H(z,w)}{\partial z\partial\bar{z}}\}_{z=w}=h_{11}(w)=\frac{\pi}{2V}\lambda(w)^{2}, (11.5)
{2H(z,w)zw}z=w=12h1(w)wh2(w),\{\frac{\partial^{2}H(z,w)}{\partial z\partial w}\}_{z=w}=\frac{1}{2}\frac{\partial h_{1}(w)}{\partial w}-h_{2}(w), (11.6)
{2H(z,w)zw¯}z=w=12h1(w)w¯h11(w).\{\frac{\partial^{2}H(z,w)}{\partial z\partial\bar{w}}\}_{z=w}=\frac{1}{2}\frac{\partial h_{1}(w)}{\partial\bar{w}}-h_{11}(w). (11.7)

The transformation properties of the coefficients hj(w)h_{j}(w) are closely related to those appearing in Lemma 11.1. Citing from [24] we have

Lemma 11.2.

Under a local holomorphic change z~=φ(z)\tilde{z}=\varphi(z) of coordinates, with w~=φ(w)\tilde{w}=\varphi(w), we have

h~0(w~)=h0(w)+Re{w~,w}0,\tilde{h}_{0}(\tilde{w})=h_{0}(w)+\operatorname{Re}\{\tilde{w},w\}_{0}, (11.8)
h~1(w~)dw~=h1(w)dw+12{w~,w}1dw,\tilde{h}_{1}(\tilde{w})d\tilde{w}=h_{1}(w)dw+\frac{1}{2}\{\tilde{w},w\}_{1}dw, (11.9)
(h~1(w~)w~2h~2(w~))dw~2=(h1(w)w2h2(w))dw2+16{w~,w}2dw2,\big{(}\frac{\partial\tilde{h}_{1}(\tilde{w})}{\partial\tilde{w}}-2\tilde{h}_{2}(\tilde{w})\big{)}d\tilde{w}^{2}=\big{(}\frac{\partial h_{1}(w)}{\partial w}-2{h}_{2}(w)\big{)}d{w}^{2}+\frac{1}{6}\{\tilde{w},w\}_{2}dw^{2}, (11.10)
h~11(w~)|dw~|2=h11(w)|dw|2.\tilde{h}_{11}(\tilde{w})|d\tilde{w}|^{2}=h_{11}(w)|dw|^{2}. (11.11)

References

  • [1] V. I. Arnold and B. A. Khesin, Topological methods in hydrodynamics, vol. 125 of Applied Mathematical Sciences, Springer-Verlag, New York, 1998.
  • [2] I. Biswas and A. K. Raina, Projective structures on a Riemann surface, Internat. Math. Res. Notices, (1996), pp. 753–768.
  • [3] S. Boatto and J. Koiller, Vortices on closed surfaces, in Geometry, mechanics, and dynamics, vol. 73 of Fields Inst. Commun., Springer, New York, 2015, pp. 185–237.
  • [4] A. Bogatskiy, Vortex flows on closed surfaces, J. Phys. A, 52 (2019), pp. 475501, 23.
  • [5] A. V. Borisov and I. S. Mamaev, On the problem of motion of vortex sources on a plane, Regul. Chaotic Dyn., 11 (2006), pp. 455–466.
  • [6] M. M. Cawte, X. Yu, B. P. Anderson, and A. S. Bradley, Snell’s law for a vortex dipole in a Bose-Einstein condensate, SciPost Phys., 6 (2019), pp. Paper No. 032, 18.
  • [7] A. J. Chorin, Numerical study of slightly viscous flow, J. Fluid Mech., 57 (1973), pp. 785–796.
  • [8] H. Cohn, Conformal mapping on Riemann surfaces, Dover Publications, Inc., New York, 1980. Reprint of the 1967 edition, Dover Books on Advanced Mathematics.
  • [9] P. J. Davis, The Schwarz function and its applications, The Mathematical Association of America, Buffalo, N. Y., 1974. The Carus Mathematical Monographs, No. 17.
  • [10] D. G. Dritschel and S. Boatto, The motion of point vortices on closed surfaces, Proc. A., 471 (2015), pp. 20140890, 25.
  • [11] B. Dubrovin, Geometry of 22D topological field theories, in Integrable systems and quantum groups (Montecatini Terme, 1993), vol. 1620 of Lecture Notes in Math., Springer, Berlin, 1996, pp. 120–348.
  • [12] H. M. Farkas and I. Kra, Riemann surfaces, vol. 71 of Graduate Texts in Mathematics, Springer-Verlag, New York, second ed., 1992.
  • [13] M. Flucher, Variational problems with concentration, vol. 36 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser Verlag, Basel, 1999.
  • [14] M. Flucher and B. Gustafsson, Vortex motion in two-dimensional hydrodynamics, Royal Instituste of Technology Research Bulletins, TRITA-MAT-1997-MA-02 (1979), pp. 1–24.
  • [15] T. Frankel, The Geometry of Physics, Cambridge University Press, Cambridge, third ed., 2012. An introduction.
  • [16] A. Fridman and P. Polubarinova, On moving singularities of a flat motion of an incompressible fluid (in Russian), Geofizicheskii Sbornik, (1928), pp. 9–23.
  • [17] P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley-Interscience [John Wiley & Sons], New York, 1978. Pure and Applied Mathematics.
  • [18] C. Grotta Ragazzo, The motion of a vortex on a closed surface of constant negative curvature, Proc. A., 473 (2017), pp. 20170447, 17.
  • [19] C. Grotta Ragazzo and H. H. d. B. Viglioni, Hydrodynamic vortex on surfaces, J. Nonlinear Sci., 27 (2017), pp. 1609–1640.
  • [20] V. Guillemin and S. Sternberg, Symplectic techniques in physics, Cambridge University Press, Cambridge, 1984.
  • [21] R. C. Gunning, Lectures on Riemann surfaces, Princeton Mathematical Notes, Princeton University Press, Princeton, N.J., 1966.
  • [22]  , Special coordinate coverings of Riemann surfaces, Math. Ann., 170 (1967), pp. 67–86.
  • [23]  , On uniformization of complex manifolds: the role of connections, vol. 22 of Mathematical Notes, Princeton University Press, Princeton, N.J., 1978.
  • [24] B. Gustafsson, Vortex motion and geometric function theory: the role of connections, Philos. Trans. Roy. Soc. A, 377 (2019), pp. 20180341, 27.
  • [25] B. Gustafsson and A. Sebbar, Critical points of Green’s function and geometric function theory, Indiana Univ. Math. J., 61 (2012), pp. 939–1017.
  • [26] D. Hally, Stability of streets of vortices on surfaces of revolution with a reflection symmetry, J. Math. Phys., 21 (1980), pp. 211–217.
  • [27] D. D. Holm and H. O. Jacobs, Multipole vortex blobs (MVB): symplectic geometry and dynamics, J. Nonlinear Sci., 27 (2017), pp. 973–1006.
  • [28] Y. Kimura, Vortex motion on surfaces with constant curvature, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci., 455 (1999), pp. 245–259.
  • [29] P. Koebe, Abhandlungen zur Theorie der konformen Abbildung, Acta Math., 41 (1916), pp. 305–344. IV. Abbildung mehrfach zusammenhängender schlichter Bereiche auf Schlitzbereiche.
  • [30] J. Koiller, Getting into the vortex: on the contributions of James Montaldi, J. Geom. Mech., 12 (2020), pp. 507–523.
  • [31] J. Koiller and S. Boatto, Vortex paires on surfaces, AIP Conference Proceedings 1130; https://doi.org/10.1063/1.3146241, 77 (2009).
  • [32] V. S. Krishnamurthy, M. H. Wheeler, D. G. Crowdy, and A. Constantin, Liouville chains: new hybrid vortex equilibria of the two-dimensional Euler equation, J. Fluid Mech., 921 (2021), pp. Paper No. A1, 35.
  • [33] K. N. Kulik, A. V. Tur, and V. V. Yanovskiĭ, Interaction of point and dipole vortices in an incompressible fluid, Teoret. Mat. Fiz., 162 (2010), pp. 459–480.
  • [34] C. C. Lin, On the motion of vortices in two dimensions. I. Existence of the Kirchhoff-Routh function, Proc. Nat. Acad. Sci. U. S. A., 27 (1941), pp. 570–575.
  • [35]  , On the motion of vortices in two dimensions. II. Some further investigations on the Kirchhoff-Routh function, Proc. Nat. Acad. Sci. U. S. A., 27 (1941), pp. 575–577.
  • [36]  , On the Motion of Vortices in Two Dimensions, University of Toronto Studies, Applied Mathematics Series, no. 5, University of Toronto Press, Toronto, Ont., 1943.
  • [37] S. G. Llewellyn Smith, How do singularities move in potential flow?, Phys. D, 240 (2011), pp. 1644–1651.
  • [38] S. G. Llewellyn Smith and R. J. Nagem, Vortex pairs and dipoles, Regul. Chaotic Dyn., 18 (2013), pp. 194–201.
  • [39] C. Marchioro and M. Pulvirenti, Mathematical theory of incompressible nonviscous fluids, vol. 96 of Applied Mathematical Sciences, Springer-Verlag, New York, 1994.
  • [40] P. K. Newton, The NN-vortex problem, vol. 145 of Applied Mathematical Sciences, Springer-Verlag, New York, 2001. Analytical techniques.
  • [41] A. Regis Rodrigues, C. Castilho, and J. Koiller, Vortex pairs on a triaxial ellipsoid and Kimura’s conjecture, J. Geom. Mech., 10 (2018), pp. 189–208.
  • [42] L. Sario and K. Oikawa, Capacity functions, Die Grundlehren der mathematischen Wissenschaften, Band 149, Springer-Verlag New York Inc., New York, 1969.
  • [43] M. Schiffer and N. S. Hawley, Connections and conformal mapping, Acta Math., 107 (1962), pp. 175–274.
  • [44] M. Schiffer and D. C. Spencer, Functionals of finite Riemann surfaces, Princeton University Press, Princeton, N. J., 1954.
  • [45] F. Schottky, Ueber die conforme Abbildung mehrfach zusammenhängender ebener Flächen, J. Reine Angew. Math., 83 (1877), pp. 300–351.
  • [46] B. F. Schutz, Jr., Geometrical methods in mathematical physics, Cambridge University Press, Cambridge, 1980.
  • [47] H. S. Shapiro, The Schwarz function and its generalization to higher dimensions, University of Arkansas Lecture Notes in the Mathematical Sciences, 9, John Wiley & Sons Inc., New York, 1992. A Wiley-Interscience Publication.
  • [48] A. Vasilev, From the Hele-Shaw experiment to integrable systems: a historical overview, Complex Anal. Oper. Theory, 3 (2009), pp. 551–585.
  • [49] Q. Wang, The nn-vortex problem on a Riemann sphere, Comm. Math. Phys., 385 (2021), pp. 565–593.