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

Spinors and horospheres

Daniel V. Mathews School of Mathematics, Monash University
[email protected]
Abstract

We give an explicit bijective correspondence between between nonzero pairs of complex numbers, which we regard as spinors or spin vectors, and horospheres in 3-dimensional hyperbolic space decorated with certain spinorial directions. This correspondence builds upon work of Penrose–Rindler and Penner. We show that the natural bilinear form on spin vectors describes a certain complex-valued distance between spin-decorated horospheres, generalising Penner’s lambda lengths to 3 dimensions.

From this, we derive several applications. We show that the complex lambda lengths in a hyperbolic ideal tetrahedron satisfy a Ptolemy equation. We also obtain correspondences between certain spaces of hyperbolic ideal polygons and certain Grassmannian spaces, under which lambda lengths correspond to Plücker coordinates, illuminating the connection between Grassmannians, hyperbolic polygons, and type A cluster algebras.

1 Introduction

Penrose and Rindler in [18] describe various aspects of relativity theory in terms of spinorial objects. The foundation of their spinorial description of spacetime is a construction associating objects of Minkowski space 1,3\mathbb{R}^{1,3} to vectors κ=(ξ,η)2\kappa=(\xi,\eta)\in\mathbb{C}^{2}, which they call spin vectors and which, for present purposes, we simply call spinors. To nonzero spinors, Penrose and Rindler associate a null flag, which is a point pp on the future light cone L+L^{+}, together with a 2-plane tangent to L+L^{+} containing pp, which behaves in a spinorial way. See Figure 1 (left).

On the other hand, Penner in [16] introduced a “decorated Teichmüller theory” which has since been highly developed (see e.g. [17]). A basic construction of this theory relates points on the future light cone L+L^{+} in Minkowski space of one lower dimension 1,2\mathbb{R}^{1,2}, to horocycles in the hyperbolic plane 2\mathbb{H}^{2}, which sits in Minkowski space as the hyperboloid model, consisting of all points 1 unit in the future of the origin. See Figure 1 (right).

L+L^{+}L+L^{+}HH2\mathbb{H}^{2}
Figure 1: Left: null flag corresponding to a spinor. Right: point on light cone (red), with corresponding horosphere HH in 2\mathbb{H}^{2} (blue).

In this paper we observe that these two constructions can be combined and generalised, yielding the following theorem.

Theorem 1.

There is an explicit, smooth, bijective, SL(2,)SL(2,\mathbb{C})-equivariant correspondence between nonzero spinors, and spin-decorated horospheres in hyperbolic 3-space 3\mathbb{H}^{3}.

A decoration on a horosphere is a tangent parallel oriented line field, i.e. a choice of direction along the horosphere at each point which is invariant under parallel translation. Such decorations exist since horospheres in 3\mathbb{H}^{3} are isometric to the Euclidean plane. A spin decoration is, roughly, a “spin lift” of such a decoration, where rotating the direction by 2π2\pi is not the identity, but rotation by 4π4\pi is; we define them precisely in Section 4. See Figure 2 (left).

The correspondence of Theorem 1 is expressed very simply in the upper half space model 𝕌\mathbb{U}. Regarding the sphere at infinity 3\partial\mathbb{H}^{3} of 3\mathbb{H}^{3} as {}\mathbb{C}\cup\{\infty\} in the usual way, the horosphere HH corresponding to (ξ,η)(\xi,\eta) has centre at ξ/η\xi/\eta. If ξ/η=\xi/\eta=\infty then HH is a horizontal plane in 𝕌\mathbb{U} at height |ξ|2|\xi|^{2}, and its decoration points in the direction iξ2i\xi^{2}. If ξ/η\xi/\eta\in\mathbb{C} then HH is a Euclidean sphere in 𝕌\mathbb{U} of Euclidean diameter |η|2|\eta|^{-2}, and its decoration at its highest point (“north pole”) points in the direction iη2i\eta^{-2}. See Figure 2 (right). We prove this in Proposition 3.9.

L+L^{+}(ξ,η)(\xi,\eta)HH3\mathbb{H}^{3} iη2i\eta^{-2}ξ/η\xi/\eta|η|2|\eta|^{-2}iξ2i\xi^{2}|ξ|2|\xi|^{2}𝕌\mathbb{U}\mathbb{C}
Figure 2: Left: Null flag corresponding to (ξ,η)(\xi,\eta), and corresponding horosphere. Right: decorated horospheres as they appear in the upper half space model 𝕌\mathbb{U}.

As indicated, both sides of this correspondence have SL(2,)SL(2,\mathbb{C})-actions. The action on 2\mathbb{C}^{2} is by the standard action of matrices on vectors. The action on horospheres is via the action of PSL(2,)PSL(2,\mathbb{C}) as orientation-preserving isometries of 3\mathbb{H}^{3}; this action lifts to SL(2,)SL(2,\mathbb{C}) to preserve spin structures.

Penrose and Rindler in [18] define a complex-valued antisymmetric bilinear form {,}\{\cdot,\cdot\} on spinors, given by the 2×22\times 2 determinant; it can also be regarded as the standard complex symplectic form on 2\mathbb{C}^{2}. On the other hand, given two horospheres H1,H2H_{1},H_{2}, we may measure the signed distance ρ\rho from one to the other along their mutually perpendicular geodesic. As detailed in Section 4, we may also measure an angle θ\theta between spin decorations, which is well defined modulo 4π4\pi. We regard d=ρ+iθd=\rho+i\theta as a complex distance between the two horospheres. See Figure 3.

Refer to caption
H1H_{1}
Refer to caption
H2H_{2}
Refer to caption
ρ\rho
θ\theta
Refer to caption
Figure 3: Complex distance between horospheres.

The correspondence of Theorem 1 extends further to relate these structures, as follows.

Theorem 2.

Given two spinors κ1,κ2\kappa_{1},\kappa_{2}, suppose the corresponding spin-decorated horospheres have complex distance dd. Then

{κ1,κ2}=exp(d2).\{\kappa_{1},\kappa_{2}\}=\exp\left(\frac{d}{2}\right).

In the 2-dimensional context, the distance between horocycles is just a real number dd, and the quantity λ=exp(d/2)\lambda=\exp(d/2) is known as a lambda length [16]. Thus, the standard bilinear form on spinors computes lambda lengths between corresponding spin-decorated horospheres.

Example.

Take κ=(1,0)\kappa=(1,0) and ω=(0,1)\omega=(0,1). In the upper half space model, κ\kappa corresponds to a horosphere centred at 1/0=1/0=\infty, so appears as a horizontal plane, at height 11 and with decoration in the direction ii. Similarly, the horosphere of ω\omega has centre 0/1=00/1=0, appearing as a sphere centred at 0, with Euclidean diameter 11, and direction at its highest point in the direction ii. These two horospheres are tangent at the point (0,0,1)(0,0,1), where their decorations align. It turns out their spin directions also align, so ρ=θ=0\rho=\theta=0 and hence λ=1={κ,ω}\lambda=1=\{\kappa,\omega\}.

Multiplying κ\kappa by a complex number reiϕre^{i\phi}, with r>0r>0 and ϕ\phi real, its horosphere remains centred at \infty, but the horizontal plane moves to height r2r^{2}, translated upwards from its original position by hyperbolic distance 2logr2\log r, and its decoration is rotated by 2ϕ2\phi. From κ\kappa to ω\omega we then have ρ=2logr\rho=2\log r and θ=2ϕ\theta=2\phi, so λ=exp(12(2logr+2ϕi))=reiϕ={κ,ω}\lambda=\exp(\frac{1}{2}(2\log r+2\phi i))=re^{i\phi}=\{\kappa,\omega\}.

Approach.

This paper proceeds through the constructions and proofs in a self-contained manner; we need to adapt and extend the constructions of Penrose–Rindler and Penner for our purposes. To give a rough overview, from a spinor κ=(ξ,η)\kappa=(\xi,\eta), following Penrose–Rindler, we obtain points (T,X,Y,Z)(T,X,Y,Z) on the positive light cone L+L^{+} in 1,3\mathbb{R}^{1,3} via

(ξη)(ξ¯η¯)=12(T+ZX+iYXiYTZ)\begin{pmatrix}\xi\\ \eta\end{pmatrix}\begin{pmatrix}\overline{\xi}&\overline{\eta}\end{pmatrix}=\frac{1}{2}\begin{pmatrix}T+Z&X+iY\\ X-iY&T-Z\end{pmatrix} (1.1)

which uses the correspondence between Hermitian 2×22\times 2 matrices and Minkowski space 1,3\mathbb{R}^{1,3} given by Pauli matrices. The flag of κ\kappa has flagpole along the ray of (T,X,Y,Z)(T,X,Y,Z) determined by (1.1), and 2-plane defined by the derivative of this map κ(T,X,Y,Z)\kappa\mapsto(T,X,Y,Z) in a certain direction depending on κ\kappa; we show this is a variation on the Penrose–Rindler construction. The correspondence between spinors and flags is SL(2,)SL(2,\mathbb{C})-equivariant, where SL(2,)SL(2,\mathbb{C}) acts on flags via its action on 1,3\mathbb{R}^{1,3} in standard fashion as SO(1,3)+SO(1,3)^{+}. In Section 2 we describe these constructions precisely, along with explicit calculations, and details of SL(2,)SL(2,\mathbb{C})-equivariance, which we have not seen proved elsewhere in the literature.

From a point pp on the figure light cone L+L^{+}, Penner’s construction in [16] is to consider the affine plane in Minkowski space consisting of all xx satisfying

x,p=1,\langle x,p\rangle=1,

where ,\langle\cdot,\cdot\rangle is the standard Lorentzian metric. This affine plane intersects the hyperboloid model of hyperbolic space \mathbb{H} in a horosphere HH. This construction works in any dimension, and in 1,3\mathbb{R}^{1,3}, we obtain an affine 3-plane which intersects the hyperboloid model of 3\mathbb{H}^{3} in a 2-dimensional horosphere. We show that when pp arises from a spinor κ\kappa, this affine 3-plane contains an affine 2-plane parallel to the flag 2-plane of κ\kappa, and this affine 2-plane intersects the horosphere in an parallel oriented line field on HH, yielding the picture of Figure 2 (left). Again, all constructions are SL(2,)SL(2,\mathbb{C})-equivariant. Precise details are given in Section 3.

Finally, all these constructions lift, in appropriate sense, to spin double covers, and remain SL(2,)SL(2,\mathbb{C})-equivariant, as we detail in Section 4.

Hyperbolic geometry applications.

These theorems have several applications and in this paper we consider some of them. In Section 5 we consider some applications to hyperbolic geometry.

Consider an ideal hyperbolic tetrahedron, i.e. with all vertices on 3\partial\mathbb{H}^{3}. Number the vertices 0,1,2,30,1,2,3. We consider a spin decoration on the tetrahedron, consisting of a spin-decorated horosphere HiH_{i} at each ideal vertex ii. There is then a complex lambda length λij\lambda_{ij} from HiH_{i} to HjH_{j}. Each λij\lambda_{ij} measures, in a certain sense, the distance between horospheres along each edge, along with the angle between them. See Figure 4.

Theorem 3.

The complex lambda lengths λij\lambda_{ij} in a spin-decorated ideal tetrahedron satisfy

λ01λ23+λ03λ12=λ02λ13.\lambda_{01}\lambda_{23}+\lambda_{03}\lambda_{12}=\lambda_{02}\lambda_{13}.

This equation is similar to Ptolemy’s theorem relating the lengths of sides and diagonals of a cyclic quadrilaterals in classical Euclidean geometry, hence we call it a Ptolemy equation. Penner in [16] proved a corresponding Ptolemy equation in 2 dimensions: when an ideal quadrilateral in 2\mathbb{H}^{2} has its vertices decorated with horocycles, the (real) lambda lengths of the edges and diagonals satisfy the same equation. Theorem 3 is a 3-dimensional generalisation showing that Ptolemy’s equation still holds, once we take horospheres to have spin decorations, and take lambda lengths to be complex. Roughly, 2-dimensional hyperbolic geometry corresponds to the case when spinors are real, i.e. lie in 2\mathbb{R}^{2}.

λ12\lambda_{12}λ01\lambda_{01}λ02\lambda_{02}λ23\lambda_{23}λ03\lambda_{03}0112233
Figure 4: Decorated horospheres and complex lambda lengths along the edges of an ideal tetrahedron.

With the previous theorems in hand, the proof of Theorem 3 is not difficult. Indeed, the four spin-decorated horospheres correspond to four spinors in 2\mathbb{C}^{2}, which can be arranged into a 2×42\times 4 matrix. The Ptolemy equation is then just the Plücker relation between 2×22\times 2 determinants of a 2×42\times 4 matrix. Alternatively, it can be seen as the relation

εABεCD+εBCεAD+εCAεBC=0\varepsilon_{AB}\varepsilon_{CD}+\varepsilon_{BC}\varepsilon_{AD}+\varepsilon_{CA}\varepsilon_{BC}=0

satisfied by the spinor ε\varepsilon, as in [18] (e.g. eq. 2.5.21).

A related result is given in Proposition 5.3, where we show that the shape parameters of an ideal tetrahedron can be recovered from these six lambda lengths.

Truncations of ideal tetrahedra along horospheres arise naturally, for instance, in complete hyperbolic structures on 3-manifolds. In a forthcoming paper with Purcell [14] we show how Ptolemy equations can be used to describe hyperbolic structures on 3-manifolds, giving a directly hyperbolic-geometric version of the Ptolemy equations described by Garoufalidis–Thurston–Zickert [10] and enhanced Ptolemy variety of Zickert [23], in turn based on work of Fock–Goncharov [4].

Even in 2 dimensions, spinors provide a useful way to analyse the geometry of horocycles; we take the spinors to have real coordinates. In forthcoming work with Zymaris we apply this to circle packing theory and generalise a classical theorem of Descartes [15].

Indeed, when ξ\xi and η\eta are both integers then the horocycles obtained in the upper half plane model of 2\mathbb{H}^{2} are the Ford circles, with their delightful relationships to Farey fractions, Diophantine approximation and continued fractions [9].

Cluster algebra applications.

In Section 6 we consider some applications to cluster algebras. We refer to [22] for an introduction to basic notions of cluster algebras.

We have already mentioned how 4 spinors arising from a spin-decorated ideal tetrahedron Δ\Delta can be arranged into a 2×42\times 4 matrix. Considering those 4 spinors up to the common action of SL(2,)SL(2,\mathbb{C}) corresponds to considering such a Δ\Delta up to isometry. And considering appropriate 2×42\times 4 matrices up to a left action by 2×22\times 2 matrices is a standard description of a Grassmannian. Thus, the correspondences of the above theorems yield a relationship between hyperbolic geometry and Grassmannians.

In [7, sec. 12.2], Fomin–Zelevinsky described a geometric realisation of cluster algebras of type AnA_{n} in terms of Grassmannians. Clusters in this case are in bijection with triangulations of an (n+3)(n+3)-gon; two clusters are joined by an edge in the exchange graph if and only if the triangulations are related by a flip [3, 8]. Fomin–Zelevinsky showed that the cluster algebra is realised by X(n+3)X(n+3), the affine cone over the Grassmannian Gr(2,n+3)\operatorname{Gr}(2,n+3) in particular, the cluster algebra there denoted 𝒜\mathcal{A}_{\circ} in type AnA_{n} is isomorphic to the \mathbb{Z}-form of the coordinate ring [X(n+3)]\mathbb{C}[X(n+3)], with the cluster variables mapping to Plücker coordinates. This has since been generalised in various ways, for instance to other Grassmannians [20] and partial flag varieties [11].

On the other hand, work of Fock–Goncharov [4], Gekhtman–Shapiro–Vainshtein [12], Fomin–Shapiro–Thurston [5] and Fomin–Thurston [6] provides geometric realisations of cluster algebras arising from surfaces, in terms of the decorated Teichmüller space T~(n+3)\widetilde{T}(n+3) introduced by Penner [16], using lambda lengths. In the particular case of an (n+3)(n+3)-gon, the lambda lengths of the diagonals provide cluster variables for the cluster algebra of type AnA_{n} [6, examples 8.10, 16.1]. Fock–Goncharov in [4] also give numerous results relating Teichmüller spaces to higher algebraic structures, but as far as they relate to hyperbolic geometry and the results of this paper, they are in dimension 2. As mentioned earlier, the Ptolemy equations of Garoufalidis–Thurston–Zickert [10], which use variables provided by Fock–Goncharov, are given a 3-dimensional hyperbolic-geometric interpretation by our complex lambda lengths in forthcoming work with Purcell [14] .

In any case, there is thus a well-understood isomorphism between the cluster algebras arising from the affine cone X(n+3)X(n+3) over the Grassmannian Gr(2,n+3)\operatorname{Gr}(2,n+3), and from the decorated Teichmüler space T~(n+3)\widetilde{T}(n+3). In [6, remark 16.2] Fomin–Thurston note a connection between the underlying spaces.

The results of this paper further illuminate the situation, by giving a direct identification of the spaces underlying these cluster algebras, and extending them to 3 dimensions. The correspondence between spinors and spin-decorated horospheres naturally yields identifications of certain decorated Teichmüller spaces, and certain Grassmannian spaces, as follows.

Theorem 4.

Let d3d\geq 3. The correspondence of Theorem 1 yields the following identifications.

  1. (i)

    The decorated Teichmüller space T~(d)\widetilde{T}(d) of ideal dd-gons is identified with the affine cone X+(n)X^{+}(n) on the positive Grassmannian Gr+(2,d)\operatorname{Gr}^{+}(2,d).

  2. (ii)

    The decorated Teichmüller space of ideal skew nn-gons in 3\mathbb{H}^{3} is identified with the affine cone X(d)X^{*}(d) on the subvariety of the complex Grassmannian Gr(2,d)\operatorname{Gr}(2,d) where all Plücker coordinates are nonzero.

Under each identification, lambda lengths correspond to Plücker coordinates.

In Section 6 we define all notions precisely and prove some properties about them, including these theorems.

The rough idea is simply that a collection of nn spinors κ1,,κn\kappa_{1},\ldots,\kappa_{n} describes the nn ideal vertices of an ideal nn-gon (in 2 dimensions), or an ideal skew nn-gon (in 3 dimensions); but on the other hand, we may place the κi\kappa_{i} as the nn columns of a 2×n2\times n matrix. The appropriate decorated Teichmüller space is then given by the certain (spin) isometry classes of such ideal (skew) nn-gons, which is an orbit space of a dd-tuple of spinors (κ1,,κd)(\kappa_{1},\ldots,\kappa_{d}). The corresponding set of orbits of 2×n2\times n matrices gives the affine cone on the appropriate Grassmannian.

It is not difficult to vary the conditions on dd-gons or Grassmannians, and find identifications between diverse versions of decorated Teichmüller spaces, and corresponding diverse Grassmannian spaces.

The appearance of the positive Grassmannian here corresponds to the fact, proved in Proposition 6.9, that a sequence of horocycles with lambda lengths that are positive, in an appropriate sense, corresponds to the the centres of the horocycles being in order around 2\partial\mathbb{H}^{2}. Positivity of determinants and Grassmannians and their relationship to ordered or convex objects arises in a similar way in the physics of scattering amplitudes, see e.g. [1].

Acknowledgments.

The author thanks Varsha for assistance in the preparation of figures. He is supported by Australian Research Council grant DP210103136.

2 Spinors to Hermitian matrices and Minkowski space

For us spinors are just elements of 2\mathbb{C}^{2}, which we regard as a complex symplectic vector space, with complex symplectic form denoted

{,}=dξdη\{\cdot,\cdot\}=d\xi\wedge d\eta

following [18]. We denote spinors by κ=(ξ,η)\kappa=(\xi,\eta) or similar. Given κ=(ξ,η)\kappa=(\xi,\eta) and κ=(ξ,η)\kappa^{\prime}=(\xi^{\prime},\eta^{\prime}) then

{κ,κ}=ξηηξ=det(ξξηη).\{\kappa,\kappa^{\prime}\}=\xi\eta^{\prime}-\eta\xi^{\prime}=\det\begin{pmatrix}\xi&\xi^{\prime}\\ \eta&\eta^{\prime}\end{pmatrix}.

We write det(κ,κ)\det(\kappa,\kappa^{\prime}) for the above determinant. We denote by 2\mathbb{C}_{*}^{2} the set of nonzero spinors.

For the purposes of linear algebra, we regard κ\kappa as a column vector and write κT\kappa^{T} for the corresponding row vector. The adjoint κ=κ¯T\kappa^{*}=\overline{\kappa}^{T} is then a row vector.

We map spinors into the set \mathcal{H} of Hermitian 2×22\times 2 matrices, or equivalently into Minkowski space 1,3\mathbb{R}^{1,3}. We take 1,3\mathbb{R}^{1,3} to have coordinates (T,X,Y,Z)(T,X,Y,Z) and metric dT2dX2dY2dZ2dT^{2}-dX^{2}-dY^{2}-dZ^{2}, denoted ,\langle\cdot,\cdot\rangle. We observe \mathcal{H} and 1,3\mathbb{R}^{1,3} are isomorphic 4-dimensional real vector spaces and we identify them in a standard way (perhaps the constant is slightly unorthodox)

(T,X,Y,Z)12[T+ZX+iYXiYTZ].(T,X,Y,Z)\leftrightarrow\frac{1}{2}\begin{bmatrix}T+Z&X+iY\\ X-iY&T-Z\end{bmatrix}.

The right hand expression is 12(T+XσX+YσY+ZσZ)\frac{1}{2}\left(T+X\sigma_{X}+Y\sigma_{Y}+Z\sigma_{Z}\right), where the σ\sigma_{\bullet} are the Pauli matrices. If a point x=(T,X,Y,Z)1,3x=(T,X,Y,Z)\in\mathbb{R}^{1,3} corresponds to SS\in\mathcal{H} then we observe TrS=T\operatorname{Tr}S=T and 4detS=x,x4\det S=\langle x,x\rangle. The light cone L={x1,3x,x=0}L=\{x\in\mathbb{R}^{1,3}\;\mid\;\langle x,x\rangle=0\} corresponds to SS with determinant zero, and the future light cone L+={xLT>0}L^{+}=\{x\in L\;\mid\;T>0\} corresponds to SS satisfying detS=0\det S=0 and TrS>0\operatorname{Tr}S>0. We define the celestial sphere 𝒮+\mathcal{S}^{+} to be the intersection of L+L^{+} with the 3-plane T=1T=1.

Definition 2.1 ([18]).

The map ϕ1\phi_{1} from 2\mathbb{C}^{2} to 1,3\mathcal{H}\cong\mathbb{R}^{1,3} is defined by ϕ1(κ)=κκ\phi_{1}(\kappa)=\kappa\kappa^{*}.

In other words,

ϕ1(κ)=(ξη)(ξ¯η¯)=(|ξ|2ξη¯ξ¯η|η|2).\phi_{1}(\kappa)=\begin{pmatrix}\xi\\ \eta\end{pmatrix}\begin{pmatrix}\overline{\xi}&\overline{\eta}\end{pmatrix}=\begin{pmatrix}|\xi|^{2}&\xi\overline{\eta}\\ \overline{\xi}\eta&|\eta|^{2}\end{pmatrix}.

We observe that the image of ϕ1\phi_{1} has determinant zero, and its diagonal entries are |ξ|2,|η|2|\xi|^{2},|\eta|^{2}, so that its trace is non-negative. Indeed it is not difficult to show ϕ1(2)=L+\phi_{1}(\mathbb{C}_{*}^{2})=L^{+}. Thus ϕ1\phi_{1} maps a 4-(real)-dimensional domain onto a 3-dimensional image. The fibres are circles; it is not difficult to show that ϕ1(κ)=ϕ1(κ)\phi_{1}(\kappa)=\phi_{1}(\kappa^{\prime}) iff κ=eiθκ\kappa=e^{i\theta}\kappa^{\prime} for some real θ\theta. Indeed, on each 3-sphere in 2\mathbb{C}^{2} given by κ\kappa with |ξ|2+|η|2|\xi|^{2}+|\eta|^{2} fixed at some constant c>0c>0, ϕ1\phi_{1} restricts to the Hopf fibration onto the 2-sphere in L+L^{+} given by T=cT=c. Thus ϕ1\phi_{1} is the cone on the Hopf fibration.

In order not to lose information, we extend ϕ1\phi_{1} to a map including tangent data. Given a tangent vector ν\nu in the real tangent space Tκ2T_{\kappa}\mathbb{C}_{*}^{2}, we write Dκϕ1(ν)D_{\kappa}\phi_{1}(\nu) for the derivative of ϕ1\phi_{1} at κ\kappa in the direction ν\nu. Since, for real tt,

ϕ1(κ+tν)=(κ+tν)(κ+tν)=κκ+(κν+νκ)t+ννt2\phi_{1}\left(\kappa+t\nu\right)=\left(\kappa+t\nu\right)\left(\kappa+t\nu\right)^{*}=\kappa\kappa^{*}+\left(\kappa\nu^{*}+\nu\kappa^{*}\right)t+\nu\nu^{*}t^{2}

we have

Dκϕ1(ν)=ddtϕ1(κ+tν)|t=0=κν+νκ.D_{\kappa}\phi_{1}(\nu)=\left.\frac{d}{dt}\phi_{1}\left(\kappa+t\nu\right)\right|_{t=0}=\kappa\nu^{*}+\nu\kappa^{*}. (2.2)

In the abstract index notation of [18], this directional derivative is κAν¯A+νAκ¯A\kappa^{A}\overline{\nu}^{A^{\prime}}+\nu^{A}\overline{\kappa}^{A^{\prime}}. At each point κ\kappa we will build a flag structure using the derivative in a certain direction Z(κ)Z(\kappa).

Definition 2.3.

The function Z:22Z\colon\mathbb{C}^{2}\longrightarrow\mathbb{C}^{2} is given by Z(ξ,η)=(iη¯,iξ¯)Z(\xi,\eta)=(i\overline{\eta},-i\overline{\xi}). In other words,

Zκ=Jκ¯whereJ=[0ii0].Z\kappa=J\overline{\kappa}\quad\text{where}\quad J=\begin{bmatrix}0&i\\ -i&0\end{bmatrix}.

Let us attempt to motivate this definition. Penrose–Rindler use a spinor τA\tau^{A} forming a spin frame, or standard symplectic basis, with κA\kappa^{A}, i.e. so that {κ,τ}=κAτA=1\{\kappa,\tau\}=\kappa_{A}\tau^{A}=1. They then form a 2-plane defined by the bivector KaLbK^{a}\wedge L^{b} where Ka=κAκ¯AK^{a}=\kappa^{A}\overline{\kappa}^{A^{\prime}} and La=κAτ¯A+τAκ¯AL^{a}=\kappa^{A}\overline{\tau}^{A^{\prime}}+\tau^{A}\overline{\kappa}^{A^{\prime}}. These two vectors are our ϕ1(κ)\phi_{1}(\kappa) and Dκϕ1(τ)D_{\kappa}\phi_{1}(\tau). But the same oriented 2-plane is obtained using any positive multiple of such τ\tau, so we could equally fix κAτA\kappa_{A}\tau^{A} simply to be positive real. Choosing τ\tau to make κAτA\kappa_{A}\tau^{A} negative real, or positive/negative imaginary, works also for our purposes. Our choice of ZZ ensures {κ,Z(κ)}=i(|ξ|2+|η|2)\{\kappa,Z(\kappa)\}=-i(|\xi|^{2}+|\eta|^{2}) is negative imaginary. Though somewhat arbitrary, this works well for our purposes.

Another perspective on ZZ is obtained by identifying (ξ,η)2(\xi,\eta)\in\mathbb{C}^{2} with the quaternion ξ+ηj\xi+\eta j. Then Zκ=kκZ\kappa=-k\kappa. On the S3S^{3} centred at the origin in 2\mathbb{C}^{2} through κ\kappa, the tangent space at κ\kappa has basis iκ,jκ,kκi\kappa,j\kappa,k\kappa. In the iκi\kappa direction lies the fibre eiθκe^{i\theta}\kappa, and ϕ1\phi_{1} is constant; ZκZ\kappa is another tangent vector to this S3S^{3}.

In any case, (2.2) and Definition 2.3 immediately yield

Dκϕ1(Zκ)=κκTJ+Jκ¯κ.D_{\kappa}\phi_{1}(Z\kappa)=\kappa\kappa^{T}J+J\overline{\kappa}\kappa^{*}. (2.4)

We now define the type of flag structure we need.

Definition 2.5.

An oriented flag of signature (d1,,dk)(d_{1},\ldots,d_{k}) in a real vector space VV is an increasing sequence of subspaces

{0}=V0V1Vk\{0\}=V_{0}\subset V_{1}\subset\cdots\subset V_{k}

where dimVi=di\dim V_{i}=d_{i}, and for i=1,,ki=1,\ldots,k, the quotient Vi/Vi1V_{i}/V_{i-1} is endowed with an orientation.

Definition 2.6.

A pointed oriented null flag, or just flag, consists of a point pL+p\in L^{+} and an oriented flag {0}V1V2\{0\}\subset V_{1}\subset V_{2} in 1,3\mathcal{H}\cong\mathbb{R}^{1,3} of signature (1,2)(1,2), such that

  1. (i)

    V1=pV_{1}=\mathbb{R}p and the orientation on V1V_{1} is towards the future (i.e. from 0 towards pp),

  2. (ii)

    V2V_{2} is a tangent plane to L+L^{+}.

The set of flags is denoted \mathcal{F}.

Thus pp is a on flagpole p\mathbb{R}p, which runs towards the future along the light cone; and the flag plane V2V_{2} is a tangent plane to the light cone, with its relative orientation equivalent to choosing the half-plane to one side of p\mathbb{R}p or the other. Note that V2V_{2} contains no timelike vectors, and p\mathbb{R}p generates the unique 1-dimensional lightlike subspace of V2V_{2}. The tangent space to L+L^{+} at pp is defined by the equation x,p=0\langle x,p\rangle=0, i.e. is the (Minkowski-)orthogonal complement pp^{\perp}. Thus pV2p\mathbb{R}p\subset V_{2}\subset p^{\perp}.

Given linearly independent pL+p\in L^{+} and vTpL+v\in T_{p}L^{+}, we denote by [[p,v]][[p,v]] the flag given by pp, the line p\mathbb{R}p oriented from the origin towards pp, the plane V2V_{2} spanned by pp and vv, and the orientation on V2/pV_{2}/\mathbb{R}p induced by vv. We observe that two flags so given [[p,v]][[p,v]], [[p,v]][[p^{\prime},v^{\prime}]] are equal if and only if p=pp=p^{\prime} and there exist real a,b,ca,b,c such that ap+bv+cv=0ap+bv+cv^{\prime}=0, where b,cb,c (which are necessarily nonzero) have opposite sign.

Note that \mathcal{F} is diffeomorphic to UTS2×UTS^{2}\times\mathbb{R}, where UTS2UTS^{2} is the unit tangent bundle of S2S^{2}: a point of S2S^{2} describes a future-oriented ray in L+L^{+}, a unit tangent vector there describes a relatively oriented 2-plane, and the \mathbb{R} factor fixes pp along the ray. Since UTS23UTS^{2}\cong\mathbb{RP}^{3} we also have 3×\mathcal{F}\cong\mathbb{RP}^{3}\times\mathbb{R}

Our version of Penrose–Rindler null flags can now be defined as the following map, upgrading ϕ1\phi_{1}.

Definition 2.7.

The map Φ1\Phi_{1} maps nonzero spinors to (pointed oriented null) flags via

Φ1:2,Φ1(κ)=[[ϕ1(κ),Dκϕ1(Zκ)]].\Phi_{1}\colon\mathbb{C}_{*}^{2}\longrightarrow\mathcal{F},\quad\Phi_{1}(\kappa)=[[\phi_{1}(\kappa),D_{\kappa}\phi_{1}(Z\kappa)]].

Thus the point ϕ1(κ)\phi_{1}(\kappa) yields the flagpole, and the derivative of ϕ1\phi_{1} in the ZκZ\kappa direction yields the relatively oriented flag plane. We verify that Dκϕ1(Zκ)D_{\kappa}\phi_{1}(Z\kappa) is (real-)linearly independent from ϕ1(κ)\phi_{1}(\kappa) using (2.4): if aκκ+b(κκTJ+Jκ¯κ)=0a\kappa\kappa^{*}+b\left(\kappa\kappa^{T}J+J\overline{\kappa}\kappa^{*}\right)=0 for some real a,ba,b then κ(aκ+bκTJ)=(Jκ¯)(bκ)\kappa\left(a\kappa^{*}+b\kappa^{T}J\right)=\left(J\overline{\kappa}\right)\left(-b\kappa^{*}\right); both sides of this equation being the product of a 2×12\times 1 and 1×21\times 2 matrix, the corresponding matrices must be proportional, say κ=cJκ¯\kappa=cJ\overline{\kappa} for some real cc; in components then ξ=ciη¯\xi=ci\overline{\eta} and η=ciξ¯\eta=-ci\overline{\xi}, so ξ=c2ξ\xi=-c^{2}\xi and η=c2η\eta=-c^{2}\eta, so that ξ=η=0\xi=\eta=0, a contradiction.

Lemma 2.8.

For two spinors κ,ν2\kappa,\nu\in\mathbb{C}_{*}^{2}, the following are equivalent:

  1. (i)

    {κ,ν}\{\kappa,\nu\} is negative imaginary (just like {κ,Zκ}\{\kappa,Z\kappa\});

  2. (ii)

    ν=aκ+bZκ\nu=a\kappa+bZ\kappa where aa is complex and bb is real positive;

  3. (iii)

    [[ϕ1(κ),Dκϕ1(ν)]]=[[ϕ1(κ),Dκϕ1(Zκ)]]=Φ1(κ).[[\phi_{1}(\kappa),D_{\kappa}\phi_{1}(\nu)]]=[[\phi_{1}(\kappa),D_{\kappa}\phi_{1}(Z\kappa)]]=\Phi_{1}(\kappa).

Proof.

If {κ,ν}\{\kappa,\nu\} is negative imaginary then {κ,bZκ}={κ,ν}\{\kappa,bZ\kappa\}=\{\kappa,\nu\} for some positive bb, and any two vectors yielding the same value for {κ,}\{\kappa,\cdot\} differ by a complex multiple of κ\kappa. This shows (i) implies (ii), and the converse is clear.

If ν=aκ+bZκ\nu=a\kappa+bZ\kappa then by linearity of the derivative, Dκϕ1(ν)=aDκϕ1(κ)+bDκϕ1(Zκ)D_{\kappa}\phi_{1}(\nu)=aD_{\kappa}\phi_{1}(\kappa)+bD_{\kappa}\phi_{1}(Z\kappa). The derivative of ϕ1\phi_{1} in the κ\kappa direction is proportional to ϕ1(κ)\phi_{1}(\kappa), and the derivative in the iκi\kappa direction is zero (pointing along a fibre of ϕ1\phi_{1}). Thus the derivatives in the ν\nu and ZκZ\kappa directions span the same plane when taken together with ϕ1(κ)\phi_{1}(\kappa); indeed, as b>0b>0, the same relatively oriented plane. In fact, this condition is equivalent to spanning the same relatively oriented plane. ∎

The spaces 2\mathbb{C}^{2}, 1,3\mathcal{H}\cong\mathbb{R}^{1,3} and \mathcal{F} all have natural SL(2,)SL(2,\mathbb{C}) actions; in all cases we denote the action of ASL(2,)A\in SL(2,\mathbb{C}) by a dot. An ASL(2,)A\in SL(2,\mathbb{C}) acts on 2\mathbb{C}^{2} by the defining representation, A.κ=AκA.\kappa=A\kappa, yielding a symplectomorphism:

{Aκ,Aκ}=det(Aκ,Aκ)=detA(κ,κ)=det(κ,κ)={κ,κ}\{A\kappa,A\kappa^{\prime}\}=\det(A\kappa,A\kappa^{\prime})=\det A(\kappa,\kappa^{\prime})=\det(\kappa,\kappa^{\prime})=\{\kappa,\kappa^{\prime}\}

since detA=1\det A=1. The same AA acts on SS\in\mathcal{H} by A.S=ASAA.S=ASA^{*}, which in 1,3\mathbb{R}^{1,3} yields in the standard way the linear maps of SO(1,3)+SO(1,3)^{+}, i.e. those which preserve the Minkowski metric and space and time orientation. The action on 1,3\mathbb{R}^{1,3} induces orientation-preserving actions on L+L^{+} and planes in 1,3\mathbb{R}^{1,3}, yielding an action on \mathcal{F}, so that A.[[p,v]]=[[A.p,A.v]]A.[[p,v]]=[[A.p,A.v]]. Essentially by definition ϕ1\phi_{1} is equivariant with respect to these actions,

ϕ1(Aκ)=AκκA=Aϕ1(κ)A=A.ϕ1(κ),\phi_{1}(A\kappa)=A\kappa\kappa^{*}A^{*}=A\phi_{1}(\kappa)A^{*}=A.\phi_{1}(\kappa),

and we have an equivariance property on its derivatives

A.Dκϕ1(ν)=DAκϕ1(Aν)A.D_{\kappa}\phi_{1}(\nu)=D_{A\kappa}\phi_{1}(A\nu)

since A.(κν+νκ)=AκνA+AνκA=(Aκ)(Aν)+(Aν)(Aκ)A.(\kappa\nu^{*}+\nu\kappa^{*})=A\kappa\nu^{*}A^{*}+A\nu\kappa^{*}A^{*}=(A\kappa)(A\nu)^{*}+(A\nu)(A\kappa)^{*}. We now show the equivariance property extends to Φ1\Phi_{1}; we have not seen a proof of this in the existing literature.

Lemma 2.9.

The map Φ1\Phi_{1} is equivariant with respect to the SL(2,)SL(2,\mathbb{C}) actions on 2\mathbb{C}_{*}^{2} and \mathcal{F}.

Proof.

We have Φ1(κ)=[[ϕ1(κ),Dκϕ1(Zκ)]]\Phi_{1}(\kappa)=[[\phi_{1}(\kappa),D_{\kappa}\phi_{1}(Z\kappa)]] so

A.Φ1(κ)=[[A.ϕ1(κ),A.Dκϕ1(Zκ)]]=[[ϕ1(Aκ),DAκϕ1(A(Zκ))]],A.\Phi_{1}(\kappa)=[[A.\phi_{1}(\kappa),A.D_{\kappa}\phi_{1}(Z\kappa)]]=[[\phi_{1}(A\kappa),D_{A\kappa}\phi_{1}(A(Z\kappa))]],

by equivariance of ϕ1\phi_{1} and its derivative. Now as AA is symplectic, {Aκ,A(Zκ)}={κ,Zκ}\{A\kappa,A(Z\kappa)\}=\{\kappa,Z\kappa\}, which is negative imaginary, so by Lemma 2.8 then [[ϕ1(Aκ),DAκϕ1(A(Zκ))]]=Φ1(Aκ)[[\phi_{1}(A\kappa),D_{A\kappa}\phi_{1}(A(Z\kappa))]]=\Phi_{1}(A\kappa). ∎

It is possible to express explicitly the linear dependence implied by the equality of the flags Φ1(Aκ)=[[ϕ1(Aκ),DAκϕ1(Z(Aκ))]]\Phi_{1}(A\kappa)=[[\phi_{1}(A\kappa),D_{A\kappa}\phi_{1}(Z(A\kappa))]] and A.Φ1(κ)=[[ϕ1(Aκ),DAκϕ1(A(Zκ))]]A.\Phi_{1}(\kappa)=[[\phi_{1}(A\kappa),D_{A\kappa}\phi_{1}(A(Z\kappa))]]: a direct computation verifies the (perhaps surprising) identity

(κTJAAκ+κAAJκ¯)ϕ1(Aκ)+(κκ)[DAκϕ1(Z(Aκ))](κAAκ)[DAκϕ1(A(Zκ))]=0.\left(\kappa^{T}JA^{*}A\kappa+\kappa^{*}A^{*}AJ\overline{\kappa}\right)\phi_{1}(A\kappa)+\left(\kappa^{*}\kappa\right)\left[D_{A\kappa}\phi_{1}\left(Z\left(A\kappa\right)\right)\right]-\left(\kappa^{*}A^{*}A\kappa\right)\left[D_{A\kappa}\phi_{1}\left(A\left(Z\kappa\right)\right)\right]=0.

We can compute Φ1\Phi_{1} completely explicitly.

Lemma 2.10.

Let κ=(ξ,η)=(a+bi,c+di)\kappa=(\xi,\eta)=(a+bi,c+di). Then in 1,3\mathbb{R}^{1,3} we have

ϕ1(κ)=(a2+b2+c2+d2,2(ac+bd),2(bcad),a2+b2c2d2),\phi_{1}(\kappa)=\left(a^{2}+b^{2}+c^{2}+d^{2},2(ac+bd),2(bc-ad),a^{2}+b^{2}-c^{2}-d^{2}\right),

and

Dκϕ1(Zκ)=(0,2(cdab),a2b2+c2d2,2(ad+bc)).D_{\kappa}\phi_{1}(Z\kappa)=\left(0,2(cd-ab),a^{2}-b^{2}+c^{2}-d^{2},2(ad+bc)\right).

The fact that Dκϕ1(Zκ)D_{\kappa}\phi_{1}(Z\kappa) has zero TT-coordinate follows from ZκZ\kappa being tangent to the S3S^{3} centred at the origin through κ\kappa, which maps under ϕ1\phi_{1} to the S2S^{2} given by the intersection of L+L^{+} with a plane at constant TT.

Proof.

This is a straightforward computation using Definition 2.1

ϕ1(κ)=[ξξ¯ξη¯ξ¯ηηη¯]=[a2+b2(ac+bd)+(bcad)i(ac+bd)(bcad)ic2+d2]\phi_{1}(\kappa)=\begin{bmatrix}\xi\overline{\xi}&\xi\overline{\eta}\\ \overline{\xi}\eta&\eta\overline{\eta}\end{bmatrix}=\begin{bmatrix}a^{2}+b^{2}&(ac+bd)+(bc-ad)i\\ (ac+bd)-(bc-ad)i&c^{2}+d^{2}\end{bmatrix}

and, via (2.4),

𝒟κϕ1(Zκ)=κκTJ+Jκ¯κ=[i(ξη¯ξη)i(ξ2+η¯2)i(ξ¯2+η2)i(ξηξη¯)].\mathcal{D}_{\kappa}\phi_{1}(Z\kappa)=\kappa\kappa^{T}J+J\overline{\kappa}\kappa^{*}=\begin{bmatrix}i\left(\overline{\xi\eta}-\xi\eta\right)&i\left(\xi^{2}+\overline{\eta}^{2}\right)\\ i\left(\overline{\xi}^{2}+\eta^{2}\right)&i\left(\xi\eta-\overline{\xi\eta}\right)\end{bmatrix}.

We denote by 3\mathbb{H}^{3} the hyperboloid model of hyperbolic 3-space

3={x=(T,X,Y,Z)1,3x,x=1,T>0}\mathbb{H}^{3}=\left\{x=(T,X,Y,Z)\in\mathbb{R}^{1,3}\;\mid\;\langle x,x\rangle=1,\;T>0\right\}

and by 3\partial\mathbb{H}^{3} the boundary at infinity of 3\mathbb{H}^{3}. So 3S2\partial\mathbb{H}^{3}\cong S^{2} and 3\partial\mathbb{H}^{3} is naturally bijective with the celestial sphere 𝒮+\mathcal{S}^{+}.

Indeed, projectivising L+L^{+} yields an the boundary at infinity 3S2\partial\mathbb{H}^{3}\cong S^{2} and under this projectivisation, 2-planes tangent to L+L^{+} containing a ray of L+L^{+} correspond bijectively with tangent lines at points of 3\partial\mathbb{H}^{3}. Moreover, relatively oriented planes containing a ray of L+L^{+} correspond bijectively with tangent directions at points of 3\partial\mathbb{H}^{3}.

The orientation-preserving isometry group SO(1,3)+SO(1,3)^{+} of 3\mathbb{H}^{3} acts transitively on the future light cone L+L^{+}, and indeed acts transitively on the tangent directions at points of 3\partial\mathbb{H}^{3}. Further, if we take an oriented flag consisting of a future-oriented line RR of L+L^{+} and a relatively oriented 2-plane π\pi tangent to L+L^{+}, then there is an element of SO(1,3)+SO(1,3)^{+} fixing RR (and its orientation) and π\pi (and its relative orientation), which sends any point on the ray to any other. Such an element is given by a hyperbolic translation along any geodesic with an endpoint at infinity corresponding to RR. In other words, SL(2,)SL(2,\mathbb{C}) acts transitively on \mathcal{F}, and the action factors through PSL(2,)SO(1,3)+PSL(2,\mathbb{C})\cong SO(1,3)^{+}.

Taking κ=(eiθ,0)\kappa=(e^{i\theta},0), we have ϕ1(κ)=(1,0,0,1)\phi_{1}(\kappa)=(1,0,0,1), which we denote p0p_{0}, and by Lemma 2.10, Φ1(κ)\Phi_{1}(\kappa) is the flag with basepoint p0p_{0} and 2-plane spanned by p0p_{0} and (0,sin2θ,cos2θ,0)(0,-\sin 2\theta,\cos 2\theta,0). Thus as we multiply κ\kappa by eiθe^{i\theta} to move through a fibre of ϕ1\phi_{1}, the flag Φ1(κ)\Phi_{1}(\kappa) rotates about a fixed pointed flagpole twice as fast. It follows that Φ1\Phi_{1} takes the value of each such flag exactly twice.

Using the equivariance of Φ1\Phi_{1} and the transitive action of SL(2,)SL(2,\mathbb{C}), the same applies for the flags based at any point on L+L^{+}. It follows that Φ1\Phi_{1} is smooth, surjective and 2–1. Moreover, the stabiliser of a flag in SO(1,3)+SO(1,3)^{+} is trivial, so that PSL(2,)PSL(2,\mathbb{C}) acts freely and transitively on \mathcal{F}. Topologically Φ1\Phi_{1} is a map 2S3×3×\mathbb{C}_{*}^{2}\cong S^{3}\times\mathbb{R}\longrightarrow\mathbb{RP}^{3}\times\mathbb{R}\cong\mathcal{F} which is a double cover.

3 From Minkowski space to horospheres

We have now built the maps ϕ1\phi_{1} and Φ1\Phi_{1} in the commutative diagram

2Φ1Φ2HorDϕ1L+ϕ2Hor\begin{array}[]{ccccc}\mathbb{C}_{*}^{2}&\stackrel{{\scriptstyle\Phi_{1}}}{{\longrightarrow}}&\mathcal{F}&\stackrel{{\scriptstyle\Phi_{2}}}{{\longrightarrow}}&\operatorname{Hor}^{D}\\ &\stackrel{{\scriptstyle\phi_{1}}}{{\searrow}}&\downarrow&&\downarrow\\ &&L^{+}&\stackrel{{\scriptstyle\phi_{2}}}{{\longrightarrow}}&\operatorname{Hor}\end{array}

where the downwards arrow L+\mathcal{F}\longrightarrow L^{+} forgets all structure of a flag except its point on L+L^{+}. In this section we define the maps ϕ2,Φ2\phi_{2},\Phi_{2}, and the spaces Hor,HorD\operatorname{Hor},\operatorname{Hor}^{D}, which involve horospheres and decorations.

Horospheres in the hyperboloid model 3\mathbb{H}^{3} are given by the intersection of 3\mathbb{H}^{3} with certain affine 3-planes in 1,3\mathbb{R}^{1,3}. Any affine 3-plane in 1,3\mathbb{R}^{1,3} is given by x1,3x\in\mathbb{R}^{1,3} satisfying an equation of the form x,n=c\langle x,n\rangle=c, where nn is a (Minkowski-)normal vector to the plane and cc is a real constant. We call such an affine 3-plane lightlike if its normal nn is lightlike. We observe that a lightlike 3-plane can be defined by an equation x,p=c\langle x,p\rangle=c where pL+p\in L^{+}; if c>0c>0 then this plane intersects 3\mathbb{H}^{3} in a horosphere, and if c0c\leq 0 the plane is disjoint from 3\mathbb{H}^{3}. Normalising such equations by requiring the constant to be 11, i.e. x,p=1\langle x,p\rangle=1, then gives a bijection between points pL+p\in L^{+} and horospheres. We denote the set of horospheres in 3\mathbb{H}^{3} by Hor\operatorname{Hor}.

Definition 3.1 ([16]).

The map ϕ2:L+Hor\phi_{2}\colon L^{+}\longrightarrow\operatorname{Hor} sends pL+p\in L^{+} to the horosphere defined by x,p=1\langle x,p\rangle=1. The map ϕ2:L+3\phi_{2}^{\partial}\colon L^{+}\longrightarrow\partial\mathbb{H}^{3} sends pp to the point at infinity of ϕ2(p)\phi_{2}(p).

Thus the map ϕ2\phi_{2} is a bijection. Indeed, it is a diffeomorphism: HorS2×\operatorname{Hor}\cong S^{2}\times\mathbb{R}, with an \mathbb{R}-family of horospheres at each point at infinity in S2=3S^{2}=\partial\mathbb{H}^{3}. Any horosphere has a unique point at infinity in 3\partial\mathbb{H}^{3}, which we also call its centre. The map ϕ2\phi_{2}^{\partial} can be regarded as the projectivisation map L+S2L^{+}\longrightarrow S^{2} or projection to the celestial sphere L+𝒮+L^{+}\longrightarrow\mathcal{S}^{+}.

Note SL(2,)SL(2,\mathbb{C}) acts naturally on 3\mathbb{H}^{3} (as on L+L^{+} and 1,3\mathbb{R}^{1,3}) in the standard way, via linear maps of SO(1,3)+SO(1,3)^{+}, and hence also on 3\partial\mathbb{H}^{3} and Hor\operatorname{Hor}, and we again denote all actions via a dot. We observe an ASL(2,)A\in SL(2,\mathbb{C}) sends the horosphere ϕ2(p)\phi_{2}(p), defined by x,p=1\langle x,p\rangle=1, to the horosphere A.ϕ2(p)A.\phi_{2}(p) defined by A1x,p=1\langle A^{-1}x,p\rangle=1. Since the action of AA preserves the Minkowski metric, this horosphere is also given by x,A.p=1\langle x,A.p\rangle=1. In other words, A.ϕ2(p)=ϕ2(A.p)A.\phi_{2}(p)=\phi_{2}(A.p) so that ϕ2\phi_{2} is SL(2,)SL(2,\mathbb{C})-equivariant. Forgetting the horospheres and recording only points at infinity, similarly ϕ2\phi_{2}^{\partial} is SL(2,)SL(2,\mathbb{C})-equivariant.

We now consider the intersection of a horosphere with a flag. So consider a horosphere ϕ2(p)\phi_{2}(p) for some pL+p\in L^{+}, and consider a flag based at the same pL+p\in L^{+}, given by the oriented sequence {0}pVp\{0\}\subset\mathbb{R}p\subset V\subset p^{\perp}. The horosphere ϕ2(p)\phi_{2}(p) is the intersection of 3\mathbb{H}^{3}, given by x,x=1\langle x,x\rangle=1, with the plane x,p=1\langle x,p\rangle=1; hence at a point qϕ2(p)q\in\phi_{2}(p), its tangent space is given by Tqϕ2(p)=pqT_{q}\phi_{2}(p)=p^{\perp}\cap q^{\perp}. The intersection of the horosphere with the flag plane VV at qq will thus be given by

Tqϕ2(p)V=qpV=qV,T_{q}\phi_{2}(p)\cap V=q^{\perp}\cap p^{\perp}\cap V=q^{\perp}\cap V,

since VpV\subset p^{\perp}. Now the intersection qVq^{\perp}\cap V is the intersection of a spacelike 3-plane q=Tq3q^{\perp}=T_{q}\mathbb{H}^{3}, and the 2-plane VV, so it is either 1- or 2-dimensional. But if it were 2-dimensional then we would have Vq=Tq3V\subset q^{\perp}=T_{q}\mathbb{H}^{3}; but VV contains a timelike vector pp, while q=Tq3q^{\perp}=T_{q}\mathbb{H}^{3} is spacelike. Thus the intersection is 1-dimensional and spacelike.

Moreover, the orientation on pV\mathbb{R}p\subset V is an orientation on V/pV/\mathbb{R}p, and thus any vector in VV not in p\mathbb{R}p obtains an orientation, depending on the side of p\mathbb{R}p to which it lies. The intersection Tqϕ2(p)V=qVT_{q}\phi_{2}(p)\cap V=q^{\perp}\cap V is spacelike, hence not equal to p\mathbb{R}p. Thus we may regard the intersection of the horosphere ϕ2(p)\phi_{2}(p) with the flag plane VV as defining an oriented line tangent to the horosphere at each point. In other words, we obtain an oriented line field on ϕ2(p)\phi_{2}(p). We denote by HorL\operatorname{Hor}^{L} the set of horospheres with oriented line fields.

Definition 3.2.

The map Φ2:HorL\Phi_{2}\colon\mathcal{F}\longrightarrow\operatorname{Hor}^{L} sends a flag {0}pV\{0\}\subset\mathbb{R}p\subset V to the horosphere ϕ2(p)\phi_{2}(p), with the oriented line field defined at each point qq by Tqϕ2(p)VT_{q}\phi_{2}(p)\cap V.

An ASL(2,)A\in SL(2,\mathbb{C}) acts on HorL\operatorname{Hor}^{L}: linear maps in SO(1,3)+SO(1,3)^{+} are orientation-preserving isometries of 3\mathbb{H}^{3}, sending horospheres to horospheres, with their derivatives sending oriented line fields to oriented line fields. Since the SL(2,)SL(2,\mathbb{C})-actions on \mathcal{F} and HorL\operatorname{Hor}^{L} are both via linear maps of SO(1,3)+SO(1,3)^{+} acting on 1,3\mathbb{R}^{1,3}, Φ2\Phi_{2} is SL(2,)SL(2,\mathbb{C})-equivariant.

It is well known that a horosphere HH is isometric to a Euclidean 2-plane. The parabolic orientation-preserving isometries of 3\mathbb{H}^{3} fixing HH act as translations on this 2-plane. This group of translations is isomorphic to the additive complex numbers. Thus, the following notion of parallelism makes sense.

Definition 3.3.

An oriented line field on a horosphere HH is parallel if it is invariant under Euclidean translations (i.e. under the action of all parabolic isometries fixing HH).

A decorated horosphere is a horosphere with a parallel oriented line field. The set of all decorated horospheres is denoted HorD\operatorname{Hor}^{D}.

Observe that to describe a parallel oriented line field on a horosphere, it suffices to give an oriented tangent line at one point; the rest of the oriented line field can then be found by parallel translation.

The following lemma calculates Φ2\Phi_{2} for a simple but useful example.

Lemma 3.4.

Φ2Φ1(1,0)\Phi_{2}\circ\Phi_{1}(1,0) is the horosphere H0H_{0} in 3\mathbb{H}^{3} which has point at infinity in the direction p0=(1,0,0,1)p_{0}=(1,0,0,1) along L+L^{+}, passing through q0=(1,0,0,0)q_{0}=(1,0,0,0), with the oriented parallel line field pointing in the direction Y=(0,0,1,0)\partial_{Y}=(0,0,1,0) at q0q_{0}.

Proof.

We have ϕ1(1,0)=p0\phi_{1}(1,0)=p_{0}, so that ϕ2ϕ1(1,0)=ϕ2(p0)\phi_{2}\circ\phi_{1}(1,0)=\phi_{2}(p_{0}) is the horosphere given by x,p0=1\langle x,p_{0}\rangle=1, which is indeed the horosphere H0H_{0}. From Lemma 2.10 the flag Φ1(1,0)\Phi_{1}(1,0) is given by [[p0,Y]][[p_{0},\partial_{Y}]], so the flag 2-plane VV is spanned by p0p_{0} and Y\partial_{Y}, with relative orientation on V/p0V/\mathbb{R}p_{0} given by Y\partial_{Y}.

Now the parabolic subgroup

P={[1c01]c}P=\left\{\begin{bmatrix}1&c\\ 0&1\end{bmatrix}\;\mid\;c\in\mathbb{C}\right\} (3.5)

fixes (1,0)(1,0) and acts simply transitively on H0H_{0}. Denoting by AcA_{c} the matrix in PP with upper right entry cc\in\mathbb{C}, the points of H0H_{0} are parametrised by cc\in\mathbb{C}; letting qc=Ac.q0q_{c}=A_{c}.q_{0} we have H0={qcc}H_{0}=\{q_{c}\;\mid\;c\in\mathbb{C}\}. We calculate the action of AcA_{c} on 1,3\mathbb{R}^{1,3} to be

Ac.(T,X,Y,Z)=(T,X,Y,Z)A_{c}.(T,X,Y,Z)=(T^{\prime},X^{\prime},Y^{\prime},Z^{\prime})

where

T=T+RecX+ImcY+|c|22(TZ),X=X+Rec(TZ),Y=Y+Imc(TZ),Z=Z+RecX+ImcY+|c|22(TZ).\begin{array}[]{cc}T^{\prime}=T+\operatorname{Re}c\,X+\operatorname{Im}c\,Y+\frac{|c|^{2}}{2}(T-Z),&X^{\prime}=X+\operatorname{Re}c\,(T-Z),\\ Y^{\prime}=Y+\operatorname{Im}c\,(T-Z),&Z^{\prime}=Z+\operatorname{Re}c\,X+\operatorname{Im}c\,Y+\frac{|c|^{2}}{2}(T-Z).\end{array}

Thus we calculate

qc=Ac.q0=(1+|c|22,Rec,Imc,|c|22)q_{c}=A_{c}.q_{0}=\left(1+\frac{|c|^{2}}{2},\;\operatorname{Re}c,\;\operatorname{Im}c,\;\frac{|c|^{2}}{2}\right)

and moreover for c,cc,c^{\prime}\in\mathbb{C} we have Ac.qc=qc+cA_{c}.q_{c^{\prime}}=q_{c+c^{\prime}}. At qcq_{c} the line field of Φ2(p0)\Phi_{2}(p_{0}) is given by qcVq_{c}^{\perp}\cap V. Now qcq_{c}^{\perp} is the 3-plane given by equation

(1+|c|22)TRecXImcY|c|22Z=0,\left(1+\frac{|c|^{2}}{2}\right)T-\operatorname{Re}c\;X-\operatorname{Im}c\;Y-\frac{|c|^{2}}{2}Z=0, (3.6)

while VV is spanned by p0p_{0} and Y\partial_{Y}, hence defined by T=ZT=Z and X=0X=0. Thus qcVq_{c}^{\perp}\cap V is defined by T=ZT=Z, X=0X=0 and Z=(Imc)YZ=(\operatorname{Im}c)\,Y, hence spanned by (Imc,0,1,Imc)=(Imc)p0+Y(\operatorname{Im}c,0,1,\operatorname{Im}c)=(\operatorname{Im}c)\,p_{0}+\partial_{Y}. Since the orientation on V/p0V/\mathbb{R}p_{0} is given by Y\partial_{Y}, the oriented line field of Φ2(p0)\Phi_{2}(p_{0}) at qcq_{c} is directed by (Imc)p0+Y(\operatorname{Im}c)\,p_{0}+\partial_{Y}. In particular, the oriented line field of Φ2(p0)\Phi_{2}(p_{0}) at q0q_{0} is directed by Y\partial_{Y}.

Now, if we apply AcA_{c} to the vector (Imc,0,1,Imc)(\operatorname{Im}c^{\prime},0,1,\operatorname{Im}c^{\prime}) directing the line field at a point qcq_{c^{\prime}} of H0H_{0}, we obtain the vector (Im(c+c),0,1,Im(c+c))(\operatorname{Im}(c+c^{\prime}),0,1,\operatorname{Im}(c+c^{\prime})) at qc+cq_{c+c^{\prime}}. Thus the oriented line field is parallel. ∎

In fact in the above calculation we observe that Ac.p0=p0A_{c}.p_{0}=p_{0} and Ac.Y=Y+Imcp0A_{c}.\partial_{Y}=\partial_{Y}+\operatorname{Im}c\,p_{0}. This shows explicitly that the parabolic subgroup PP preserves the flag plane VV, and in fact acts as the identity on both p0\mathbb{R}p_{0} and V/p0V/\mathbb{R}p_{0}.

In fact this example is generic enough to give the following.

Lemma 3.7.

The map Φ2\Phi_{2} is a diffeomorphism HorD\mathcal{F}\longrightarrow\operatorname{Hor}^{D}.

In other words, the oriented line field of any flag is parallel, and Φ2\Phi_{2} provides a smooth correspondence between flags and decorated horospheres.

Proof.

First we show Φ2\Phi_{2} always yields parallel oriented line fields. Lemma 3.4 shows this when Φ2\Phi_{2} is applied to Φ1(1,0)\Phi_{1}(1,0)\in\mathcal{F}. But the action of SL(2,)SL(2,\mathbb{C}) is transitive on \mathcal{F}, and the action of SL(2,)SL(2,\mathbb{C}) on 3\mathbb{H}^{3} (hence on horospheres) is by isometries, and Φ2\Phi_{2} is SL(2,)SL(2,\mathbb{C})-equivariant. Any flag in \mathcal{F} is thus of the form A.Φ1(1,0)A.\Phi_{1}(1,0) for some ASL(2,)A\in SL(2,\mathbb{C}), so Φ2A.Φ1(1,0)=A.Φ2Φ1(1,0)\Phi_{2}A.\Phi_{1}(1,0)=A.\Phi_{2}\Phi_{1}(1,0), which has parallel oriented line field.

Next we show that Φ2\Phi_{2} sends the flags of the form [[p0,v]][[p_{0},v]] bijectively to decorations on H0H_{0}. These flags are those of the form

Φ1(eiθ,0)=[eiθ00eiθ].Φ1(1,0)=[[p0,(0,sin2θ,cos2θ,0)]],\Phi_{1}(e^{i\theta},0)=\begin{bmatrix}e^{i\theta}&0\\ 0&e^{-i\theta}\end{bmatrix}.\Phi_{1}(1,0)=[[p_{0},(0,-\sin 2\theta,\cos 2\theta,0)]],

as calculated above. Denote the latter vector by θ\partial_{\theta}, so Φ1(eiθ,0)=[[p0,θ]]\Phi_{1}(e^{i\theta},0)=[[p_{0},\partial_{\theta}]]. Then Φ2Φ1(eiθ,0)\Phi_{2}\Phi_{1}(e^{i\theta},0) is the horosphere H0H_{0}, with oriented parallel line field at q0q_{0} given by the intersection of the flag 2-plane with q0q_{0}^{\perp}. Since q0q_{0}^{\perp} is given by T=0T=0 (equation ((3.6)) with c=0c=0), which contains θ\partial_{\theta}, the oriented line field of Φ2Φ1(eiθ,0)\Phi_{2}\Phi_{1}(e^{i\theta},0) at q0q_{0} is directed by θ\partial_{\theta}. As θ\theta increases say from 0 to π\pi, both the flag through p0p_{0} and the decoration on H0H_{0} rotate through a full 2π2\pi, with Φ2\Phi_{2} providing a bijection.

We have already seen that ϕ2\phi_{2} provides a bijection between L+L^{+} and Hor\operatorname{Hor}; using the transitivity of SL(2,)SL(2,\mathbb{C}) on \mathcal{F} and HorD\operatorname{Hor}^{D}, and equivariance of Φ2\Phi_{2}, it follows that Φ2\Phi_{2} provides a bijection between the flags based at each p0L+p_{0}\in L^{+}, and decorations on the corresponding horospheres ϕ2(p0)\phi_{2}(p_{0}).

Thus Φ2\Phi_{2} is a bijection. It and its inverse are clearly smooth, once \mathcal{F} and HorD\operatorname{Hor}^{D} are given their natural smooth structures. We have already seen UTS2×\mathcal{F}\cong UTS^{2}\times\mathbb{R}. The space of horospheres is naturally HorS2×\operatorname{Hor}\cong S^{2}\times\mathbb{R}, and decorations can be given by unit tangent vectors to the sphere at infinity, so that HorDUTS2×\operatorname{Hor}^{D}\cong UTS^{2}\times\mathbb{R}. ∎

We now consider our horospheres in the upper half space model 𝕌\mathbb{U} of 3\mathbb{H}^{3}, given in the usual way as

𝕌={(x,y,z)3z>0}with metricds2=dx2+dy2+dz2z2.\mathbb{U}=\left\{(x,y,z)\in\mathbb{R}^{3}\;\mid\;z>0\right\}\quad\text{with metric}\quad ds^{2}=\frac{dx^{2}+dy^{2}+dz^{2}}{z^{2}}.

As usual we identify the plane z=0z=0 with \mathbb{C} and 𝕌\partial\mathbb{U} with \mathbb{C}\cup\infty, and coordinates (x,y)(x,y) with x+yix+yi\in\mathbb{C}. In 𝕌\mathbb{U}, horospheres centred at \infty appear as horizontal planes; we call the zz-coordinate of this plane the height of the horosphere. Horospheres centred at other points appear as spheres tangent to \mathbb{C}; we call the maximum of zz on the sphere the Euclidean diameter of the horosphere.

We proceed from 3\mathbb{H}^{3} to 𝕌\mathbb{U} via the disc model 𝔻\mathbb{D}. We have the standard isometries given by

3𝔻,(T,X,Y,Z)11+T(X,Y,Z)and𝔻𝕌,(x,y,z)x+iy1z,\mathbb{H}^{3}\longrightarrow\mathbb{D},\quad(T,X,Y,Z)\mapsto\frac{1}{1+T}\left(X,Y,Z\right)\quad\text{and}\quad\partial\mathbb{D}\longrightarrow\partial\mathbb{U},\quad(x,y,z)\mapsto\frac{x+iy}{1-z}, (3.8)

where in the latter map we regard 𝔻\partial\mathbb{D} as the standard S23S^{2}\subset\mathbb{R}^{3} and 𝕌\partial\mathbb{U} as {}\mathbb{C}\cup\{\infty\}. Of course SL(2,)SL(2,\mathbb{C})-actions carry through equivariantly to each model as isometries, and on 𝕌\partial\mathbb{U}\cong\mathbb{C}\cup\infty the action is via Möbius transformations in the usual way,

[αβγδ].z=αz+βγz+δ.\begin{bmatrix}\alpha&\beta\\ \gamma&\delta\end{bmatrix}.z=\frac{\alpha z+\beta}{\gamma z+\delta}.

We now introduce some terminology to describe decorations, i.e. parallel oriented line fields, on horospheres in 𝕌\mathbb{U}. A horosphere centred at \infty is a horizontal plane parallel to \mathbb{C}, so a parallel oriented line field appears as a line field invariant under Euclidean translations, and can be described by a complex number which points in the direction of the lines. This complex number is well defined up to positive multiples and we say it specifies the decoration. On a horosphere centred elsewhere, we can describe an oriented line field by giving a vector directing it at the point with maximum zz-coordinate (its “north pole”); since the tangent plane there is also parallel to \mathbb{C}, we can also describe it by a complex number, up to positive multiple. We call this a north pole specification of a decoration.

We can now give the decorated horospheres corresponding to spinors explicitly, verifying the description in the introduction, illustrated in Figure 2 (right).

Proposition 3.9.

The spinor (ξ,η)(\xi,\eta) maps under Φ2Φ1\Phi_{2}\circ\Phi_{1} to a decorated horosphere whose centre is at ξ/η\xi/\eta in the upper half space model.

  1. (i)

    If η0\eta\neq 0 then the horosphere has Euclidean diameter |η|2|\eta|^{-2}, and decoration north-pole specified by i/η2i/\eta^{2},

  2. (ii)

    If η=0\eta=0, then the horosphere has height |ξ|2|\xi|^{2}, and decoration specified by iξ2i\xi^{2}.

In particular, forgetting the decorations, the above proposition gives an explicit description of ϕ2ϕ1(ξ,η)\phi_{2}\circ\phi_{1}(\xi,\eta). And forgetting all but the centres of the horospheres, it yields ϕ2ϕ1(ξ,η)=ξη\phi_{2}^{\partial}\circ\phi_{1}(\xi,\eta)=\frac{\xi}{\eta}.

Proof.

Letting ξ=a+bi\xi=a+bi, η=c+di\eta=c+di, ϕ1(ξ,η)\phi_{1}(\xi,\eta) is given in Lemma 2.10. . Then ϕ2\phi_{2}^{\partial}, for the hyperboloid model, just projectivises the rays of L+L^{+} to points; taking (X,Y,Z)(X,Y,Z) for the point on each ray with T=1T=1 gives ϕ2ϕ1(ξ,η)\phi_{2}^{\partial}\phi_{1}(\xi,\eta) on 𝔻\partial\mathbb{D} as

1a2+b2+c2+d2(2(ac+bd), 2(bcad),a2+b2c2d2).\frac{1}{a^{2}+b^{2}+c^{2}+d^{2}}\left(2(ac+bd),\,2(bc-ad),\,a^{2}+b^{2}-c^{2}-d^{2}\right).

The centre of the horosphere on 𝕌\partial\mathbb{U}\cong\mathbb{C}\cup\infty is then, using (3.8),

2(ac+bd)a2+b2+c2+d2+i2(bcad)a2+b2+c2+d21a2+b2c2d2a2+b2+c2+d2=a+bic+di=ξη.\frac{\frac{2(ac+bd)}{a^{2}+b^{2}+c^{2}+d^{2}}+i\frac{2(bc-ad)}{a^{2}+b^{2}+c^{2}+d^{2}}}{1-\frac{a^{2}+b^{2}-c^{2}-d^{2}}{a^{2}+b^{2}+c^{2}+d^{2}}}=\frac{a+bi}{c+di}=\frac{\xi}{\eta}.

From Lemma 3.4, Φ2Φ1(1,0)\Phi_{2}\circ\Phi_{1}(1,0), in the hyperboloid model, is the horosphere centred at p0=(1,0,0,1)p_{0}=(1,0,0,1), passing through q0=(1,0,0,0)q_{0}=(1,0,0,0), and at q0q_{0} has decoration in the direction Y=(0,0,1,0)\partial_{Y}=(0,0,1,0). In 𝔻\mathbb{D}, this corresponds to the horosphere centred at (0,0,1)(0,0,1), passing through (0,0,0)(0,0,0), and having decoration in the direction (0,1,0)(0,1,0) there. In 𝕌\mathbb{U}, this corresponds to the horosphere centred at \infty, passing through (0,0,1)(0,0,1), and having decoration in the direction (0,1,0)(0,1,0) at that point. In other words, it has height 11 and decoration specified by ii.

The decorated horospheres Φ2Φ1(ξ,η)\Phi_{2}\circ\Phi_{1}(\xi,\eta) can now be found in general using SL(2,)SL(2,\mathbb{C})-equivariance. Observe that

(01)=[0110].(10),(ξ0)=[ξ00ξ1].(10),(ξη)=[η1ξ0η].(01)\begin{pmatrix}0\\ 1\end{pmatrix}=\begin{bmatrix}0&-1\\ 1&0\end{bmatrix}.\begin{pmatrix}1\\ 0\end{pmatrix},\quad\begin{pmatrix}\xi\\ 0\end{pmatrix}=\begin{bmatrix}\xi&0\\ 0&\xi^{-1}\end{bmatrix}.\begin{pmatrix}1\\ 0\end{pmatrix},\quad\begin{pmatrix}\xi\\ \eta\end{pmatrix}=\begin{bmatrix}\eta^{-1}&\xi\\ 0&\eta\end{bmatrix}.\begin{pmatrix}0\\ 1\end{pmatrix} (3.10)

Thus the decorated horosphere of (0,1)(0,1) is obtained from the decorated horosphere of (1,0)(1,0) by applying the Möbius transformation z1zz\mapsto\frac{-1}{z}; hence it is centred at 0, has Euclidean diameter 11, and is north-pole specified by ii. Similarly, the decorated horosphere of (ξ,0)(\xi,0), for ξ0\xi\neq 0, is obtained from that of (1,0)(1,0) by applying zξ2zz\mapsto\xi^{2}z, hence is centred at \infty, has height |ξ|2|\xi|^{2}, and is specified by iξ2i\xi^{2}. And the decorated horosphere of (ξ,η)(\xi,\eta), for η0\eta\neq 0, is obtained from that of (0,1)(0,1) by applying zη2z+ξηz\mapsto\eta^{-2}z+\frac{\xi}{\eta}, hence is centred at ξ/η\xi/\eta, has Euclidean diameter |η|2|\eta|^{-2}, and is north-pole specified by iη2i\eta^{-2}. ∎

Thus, if we multiply a spinor κ\kappa by a complex number reiθre^{i\theta}, with r>0r>0 and θ\theta real, the effect on the corresponding horosphere HH is to translate it by distance 2logr2\log r along any geodesic γ\gamma perpendicular to HH oriented towards its centre, and rotate the decoration by 2θ2\theta about γ\gamma.

4 Spin decorations and complex lambda lengths

We now introduce the concepts necessary to explain the lifts of previous constructions to spin double covers, and the notion of complex lambda length between two spin-decorated horospheres. In this section 3\mathbb{H}^{3} refers to hyperbolic 3-space, regardless of model.

We use the cross product ×\times in 3\mathbb{H}^{3} in the elementary sense that if v,wv,w are tangent to 3\mathbb{H}^{3} at a common point pp, making an angle of θ\theta, then v×wv\times w is tangent to 3\mathbb{H}^{3} at pp, has length |v||w|sinθ|v|\,|w|\sin\theta, and points in the direction perpendicular to vv and ww given by the right-hand rule.

A horosphere HH in 3\mathbb{H}^{3} (like any oriented surface in a 3-manifold) has two normal directions: we call the direction towards its centre outward (“pointing out of 3\mathbb{H}^{3}”), and the direction away from its centre inward (“pointing into 3\mathbb{H}^{3}”). There are well-defined outward and inward unit normal vector fields along HH, which we denote Nout,NinN^{out},N^{in} respectively.

By a frame we mean a right-handed orthonormal frame at a point in 3\mathbb{H}^{3}, i.e. a triple of orthogonal unit vectors (f1,f2,f3)(f_{1},f_{2},f_{3}) such that f1×f2=f3f_{1}\times f_{2}=f_{3}. The collection of frames then forms a principal SO(3)SO(3)-bundle over 3\mathbb{H}^{3} which we denote

Fr3.\operatorname{Fr}\longrightarrow\mathbb{H}^{3}.

We may take its spin double (universal) cover, which we denote

FrS3,\operatorname{Fr}^{S}\longrightarrow\mathbb{H}^{3},

which is a principal Spin(3)\operatorname{Spin}(3)-bundle. We refer to points of FrS\operatorname{Fr}^{S} as spin frames. Each point in Fr\operatorname{Fr} has two lifts in FrS\operatorname{Fr}^{S}, i.e. each frame lifts to two spin frames.

The group of orientation-preserving symmetries of 3\mathbb{H}^{3} is naturally isomorphic to PSL(2,)PSL(2,\mathbb{C}), and acts simply transitively on Fr\operatorname{Fr}. Choosing a basepoint F0F_{0} in Fr\operatorname{Fr} then we may obtain an explicit identification PSL(2,)FrPSL(2,\mathbb{C})\cong\operatorname{Fr}, given by MM.F0M\leftrightarrow M.F_{0} for MPSL(2,)M\in PSL(2,\mathbb{C}).

Similarly, SL(2,)SL(2,\mathbb{C}) acts simply transitively on FrS\operatorname{Fr}^{S}. And the identification PSL(2,)FrPSL(2,\mathbb{C})\cong\operatorname{Fr} lifts to double covers, after we choose a lifted basepoint F0~\widetilde{F_{0}}, giving an explicit diffeomorphism SL(2,)FrSSL(2,\mathbb{C})\cong\operatorname{Fr}^{S} as AA.F0~A\leftrightarrow A.\widetilde{F_{0}}. The two matrices A,ASL(2,)A,-A\in SL(2,\mathbb{C}) lifting ±APSL(2,)\pm A\in PSL(2,\mathbb{C}) then correspond to the two spin frames A.F0A.F_{0}, A.F0-A.F_{0} lifting the frame (±A).F0(\pm A).F_{0}. These two spin frames are related by a 2π2\pi rotation about any axis at their common point. We can regard elements of SL(2,)SL(2,\mathbb{C}) as spin isometries; each isometry in PSL(2,)PSL(2,\mathbb{C}) lifts to two spin isometries, which differ by a 2π2\pi rotation. Since SL(2,)SL(2,\mathbb{C}) is the universal cover of the isometry group PSL(2,)PSL(2,\mathbb{C}), we can also regard elements of SL(2,)SL(2,\mathbb{C}) as homotopy classes of paths of isometries starting at the identity.

From a decoration on a horosphere HH, normalised to a parallel unit tangent vector field vv on HH, we can then construct frame fields along HH as follows.

Definition 4.1.

Let vv be a unit parallel tangent vector field on a horosphere HH.

  1. (i)

    The inward frame field of vv is the frame field on HH given by Fin=(Nin,v,Nin×v)F^{in}=(N^{in},v,N^{in}\times v).

  2. (ii)

    The outward frame field of vv is the frame field on HH given by Fout=(Nout,v,Nout×v)F^{out}=(N^{out},v,N^{out}\times v).

Indeed a decorated horosphere is uniquely specified by its inward and outward frame fields and so we can denote a decorated horosphere by (H,F)(H,F) where FF is the pair of frames F=(Fin,Fout)F=(F^{in},F^{out}).

A frame field is a continuous section of Fr\operatorname{Fr} along HH, and it has two lifts to Spin\operatorname{Spin}.

Definition 4.2.

An outward (resp. inward) spin decoration on HH is a continuous lift of an outward (resp. inward) frame field from Fr\operatorname{Fr} to FrS\operatorname{Fr}^{S}.

From the inward frame field (Nin,v,Nin×v)(N^{in},v,N^{in}\times v) of a unit parallel vector field vv on HH, one can rotate the frame at each point of HH by an angle of π\pi or π-\pi about vv to obtain the outward frame field of vv, and vice versa. After taking an inward spin decoration lifting the inward frame field, one can similarly rotate the frame at each point by an angle of π\pi about vv, which will result in an outward spin decoration. However, rotations of π\pi or π-\pi about vv yield distinct results, related by a 2π2\pi rotation. Thus we make the following definition, which is a somewhat arbitrary convention, but we need it for our results to hold.

Definition 4.3.
  1. (i)

    Let WoutW^{out} be an outward spin decoration on HH. The associated inward spin decoration is the spin decoration obtained by rotating WoutW^{out} by angle π\pi about vv at each point of HH.

  2. (ii)

    Let WinW^{in} be an inward spin decoration on HH. The associated outward spin decoration is the spin decoration obtained by rotating WinW^{in} by angle π-\pi about vv at each point of HH.

We observe that associated spin decorations come in pairs W=(Win,Wout)W=(W^{in},W^{out}), each associated to the other.

Definition 4.4.

A spin decoration on a horosphere HH is a pair W=(Win,Wout)W=(W^{in},W^{out}) of associated inward and outward spin decorations. We denote a spin-decorated horosphere by (H,W)(H,W), and denote the set of spin-decorated horospheres by HorS\operatorname{Hor}^{S}.

Note that under the identification PSL(2,)FrPSL(2,\mathbb{C})\cong\operatorname{Fr}, with an appropriate choice of basepoint frame, the parabolic subgroup PP of equation (3.5) (or more precisely its image ±P\pm P in PSL(2,)PSL(2,\mathbb{C})) corresponds to all the frames of the outward frame field of Φ2Φ1(1,0)\Phi_{2}\circ\Phi_{1}(1,0). The cosets of ±P\pm P then correspond bijectively with decorated horospheres. Similarly, under the identification SL(2,)FrSSL(2,\mathbb{C})\cong\operatorname{Fr}^{S} with an appropriate choice of basepoint, the cosets of PP correspond bijectively with spin-decorated horospheres:

PSL(2,)/(±P)HorD,SL(2,)/PHorS.PSL(2,\mathbb{C})/(\pm P)\cong\operatorname{Hor}^{D},\quad SL(2,\mathbb{C})/P\cong\operatorname{Hor}^{S}.

We now consider lifts of the maps

2Φ1Φ2HorD\mathbb{C}_{*}^{2}\stackrel{{\scriptstyle\Phi_{1}}}{{\longrightarrow}}\mathcal{F}\stackrel{{\scriptstyle\Phi_{2}}}{{\longrightarrow}}\operatorname{Hor}^{D}

Topologically, we have 2S3×\mathbb{C}_{*}^{2}\cong S^{3}\times\mathbb{R}, we have seen HorDUTS2×3×\mathcal{F}\cong\operatorname{Hor}^{D}\cong UTS^{2}\times\mathbb{R}\cong\mathbb{RP}^{3}\times\mathbb{R}, and we have seen Φ1\Phi_{1} is a double cover and Φ2\Phi_{2} is a diffeomorphism. Indeed, all the spaces here are S1×S^{1}\times\mathbb{R}\cong\mathbb{C}_{*} bundles over S2S^{2} and the maps are bundle maps, which in an appropriate sense are the identity on the base space S2S^{2}. The spaces \mathcal{F} and HorD\operatorname{Hor}^{D} both have fundamental group /2\mathbb{Z}/2, and we can consider their double (hence universal) covers. A nontrivial loop in \mathcal{F} is given by fixing a flagpole and rotating a flag through 2π2\pi; in the double cover, rotating the flag through 2π2\pi is no longer a loop, but rotating the flag through 4π4\pi gives a loop.

Definition 4.5.

The double cover of the space of flags \mathcal{F} is denoted S\mathcal{F}^{S}. We call its elements spin flags.

Our spin flags are the null flags of [18].

A nontrivial loop in HorD\operatorname{Hor}^{D} is given by fixing a horosphere and rotating its decoration through 2π2\pi. In the double cover, a rotation through 2π2\pi is not a loop but a rotation through 4π4\pi gives a loop. In other words, the double cover of HorD\operatorname{Hor}^{D} is HorS\operatorname{Hor}^{S}. Choosing basepoints (arbitrarily) one then obtains lifts Φ1~,Φ2~\widetilde{\Phi_{1}},\widetilde{\Phi_{2}} such that the diagram

C2Φ1~SΦ2~HorSΦ1Φ2HorD\begin{array}[]{ccccc}C_{*}^{2}&\stackrel{{\scriptstyle\widetilde{\Phi_{1}}}}{{\longrightarrow}}&\mathcal{F}^{S}&\stackrel{{\scriptstyle\widetilde{\Phi_{2}}}}{{\longrightarrow}}&\operatorname{Hor}^{S}\\ &\stackrel{{\scriptstyle\Phi_{1}}}{{\searrow}}&\downarrow&&\downarrow\\ &&\mathcal{F}&\stackrel{{\scriptstyle\Phi_{2}}}{{\longrightarrow}}&\operatorname{Hor}^{D}\end{array}

commutes, where the downwards arrows are double covering maps. The action of SL(2,)SL(2,\mathbb{C}) (which is simply connected) lifts to actions on these covers and all maps remain SL(2,)SL(2,\mathbb{C})-equivariant.

Thus a spinor κ\kappa maps under Φ2~Φ1~\widetilde{\Phi_{2}}\circ\widetilde{\Phi_{1}} to a spin-decorated horosphere lifting the decorated horosphere described in Proposition 3.9. Multiplying κ\kappa by reiθre^{i\theta}, with r>0r>0 and θ\theta real, still translates it 2logr2\log r towards its centre and rotates the decoration by 2θ2\theta, but now the rotation is taken modulo 4π4\pi.

We can now prove Theorem 1, that there is an explicit smooth bijective correspondence between 2\mathbb{C}_{*}^{2} and HorS\operatorname{Hor}^{S}.

Proof of Theorem 1.

At the end of Section 2 we observed that Φ1\Phi_{1} is a smooth double cover, topologically S3×3S^{3}\times\mathbb{R}\longrightarrow\mathbb{RP}^{3}\longrightarrow\mathbb{R}. In Lemma 3.7 we showed Φ2\Phi_{2} is a diffeomorphism. Their lifts Φ1~\widetilde{\Phi_{1}} and Φ2~\widetilde{\Phi_{2}} are then both diffeomorphisms, topologically S3×S3×S^{3}\times\mathbb{R}\longrightarrow S^{3}\times\mathbb{R}. We have defined these maps explicitly. We have also shown all maps are SL(2,)SL(2,\mathbb{C})-equivariant. Thus Φ2~Φ1~\widetilde{\Phi_{2}}\circ\widetilde{\Phi_{1}} provides the claimed correspondence. ∎

We use spin frames to define complex lambda lengths between spin-decorated horospheres. For this, we need to compare frames along geodesics, and we need frames to be adapted to geodesics, in a suitable sense. (Here, as throughout, frames are right-handed and orthonormal.)

Definition 4.6.

Let pp be a point on an oriented geodesic γ\gamma in 3\mathbb{H}^{3}. A frame F=(f1,f2,f3)F=(f_{1},f_{2},f_{3}) at pp is adapted to γ\gamma if f1f_{1} is positively tangent to γ\gamma. A spin frame F~\widetilde{F} at pp is adapted to γ\gamma if it is the lift of a frame adapted to γ\gamma.

Now if we have two points p1,p2p_{1},p_{2} on an oriented geodesic γ\gamma, and frames Fi=(f1i,f2i,f3i)F^{i}=(f_{1}^{i},f_{2}^{i},f_{3}^{i}) at each pip_{i}, adapted to γ\gamma, then there is then a screw motion along γ\gamma which takes F1F^{1} to F2F^{2} as follows. Being adapted to γ\gamma, the first vectors f11f_{1}^{1} and f12f_{1}^{2} in each frame point along γ\gamma. Parallel translation along γ\gamma from p1p_{1} to p2p_{2} takes F1F^{1} to a frame F1F^{\prime 1} at p2p_{2} which agrees with F2F^{2} in its first vector. This translation is by a signed distance ρ\rho which we regard as positive or negative according to the orientation on γ\gamma. A further rotation of some angle θ\theta about γ\gamma (signed using the orientation of γ\gamma) then moves F1F^{\prime 1} to F2F^{2}. Note that θ\theta is only well defined modulo 2π2\pi. However we may repeat this process with spin frames, and then θ\theta is well defined modulo 4π4\pi.

Definition 4.7.

Let F1,F2F^{1},F^{2} be frames, or spin frames, at points p1,p2p_{1},p_{2} on an oriented geodesic γ\gamma, adapted to γ\gamma. The complex distance from F1F^{1} to F2F^{2} is ρ+iθ\rho+i\theta, where a translation along γ\gamma of signed distance ρ\rho, followed by a rotation about γ\gamma of angle θ\theta, takes F1F^{1} to F2F^{2}.

In general two frames are not adapted to a common oriented geodesic, but when two frames are adapted to a common oriented geodesic, that oriented geodesic is unique, and so we may speak of the complex distance between the frames. The same applies to spin frames. Note that the complex distance between frames adapted to a common geodesic is well defined modulo 2πi2\pi i; between spin frames, it is well defined modulo 4πi4\pi i.

We can now define complex lambda lengths between decorated and spin-decorated horospheres. Let H1,H2H_{1},H_{2} be horospheres, let zi3z_{i}\in\partial\mathbb{H}^{3} be the centre of HiH_{i}, and let γij\gamma_{ij} be the oriented geodesic from ziz_{i} to zjz_{j}. Thus γ12\gamma_{12} and γ21\gamma_{21} are the two orientations of the unique common perpendicular to the horospheres. Let pi=γ12Hip_{i}=\gamma_{12}\cap H_{i}. If the HiH_{i} are decorated, we have pairs Fi=(Fiin,Fiout)F_{i}=(F^{in}_{i},F^{out}_{i}) of inward and outward frame fields on each HiH_{i}, and note that F1in(p1)F^{in}_{1}(p_{1}) and F2out(p2)F^{out}_{2}(p_{2}) are both adapted to γ12\gamma_{12}. If the HiH_{i} are spin-decorated, we have pairs Wi=(Wiin,Wiout)W_{i}=(W^{in}_{i},W^{out}_{i}) of associated inward and outward spin decorations on each HiH_{i}, and we note that W1in(p1)W^{in}_{1}(p_{1}) and W2out(p2)W^{out}_{2}(p_{2}) are adapted to γ12\gamma_{12}.

Definition 4.8.
  1. (i)

    If (H1,F1)(H_{1},F_{1}) and (H2,F2)(H_{2},F_{2}) are decorated horospheres, the complex lambda length from (H1,F1)(H_{1},F_{1}) to (H2,F2)(H_{2},F_{2}) is

    λ12=exp(d2),\lambda_{12}=\exp\left(\frac{d}{2}\right),

    where dd is the complex distance from F1in(p1)F^{in}_{1}(p_{1}) to F2out(p2)F^{out}_{2}(p_{2}).

  2. (ii)

    If (H1,W1)(H_{1},W_{1}) and (H2,W2)(H_{2},W_{2}) are spin-decorated horospheres, the complex lambda length from (H1,W1)(H_{1},W_{1}) to (H2,W2)(H_{2},W_{2}) is

    λ12=exp(d2),\lambda_{12}=\exp\left(\frac{d}{2}\right),

    where dd is the complex distance from Win(p1)W^{in}(p_{1}) to Wout(p2)W^{out}(p_{2}).

When the horospheres H1H_{1} and H2H_{2} have a common centre, then the complex lambda length between them is zero in either case.

Note that for decorated horospheres, dd is only well defined modulo 2πi2\pi i, so λ12\lambda_{12} is only well defined up to sign. For spin-decorated horospheres however dd is well defined modulo 4πi4\pi i, so λ12\lambda_{12} is a well defined complex number, and indeed we have a well defined function λ:HorS×HorS\lambda\colon\operatorname{Hor}^{S}\times\operatorname{Hor}^{S}\longrightarrow\mathbb{C}.

We observe that λ\lambda is in fact continuous. In particular, if two horospheres move so that their centres approach each other, then the length of the segment of their common perpendicular geodesic which lies in the intersection of the horoballs becomes arbitrarily large, so Red\operatorname{Re}d\rightarrow-\infty and hence λ0\lambda\rightarrow 0.

In fact, as we now see, λ\lambda is antisymmetric.

Lemma 4.9.

Let (H1,W1)(H_{1},W_{1}), (H2,W2)(H_{2},W_{2}) be spin-decorated horospheres, and let λij\lambda_{ij} be the complex lambda length from (Hi,Wi)(H_{i},W_{i}) to (Hj,Wj)(H_{j},W_{j}). Then λ12=λ21\lambda_{12}=-\lambda_{21}.

Proof.

If H1,H2H_{1},H_{2} have common centre λ12=λ21=0\lambda_{12}=\lambda_{21}=0. So we may assume H1,H2H_{1},H_{2} have distinct centres z1,z2z_{1},z_{2}. As above, let γij\gamma_{ij} be the oriented geodesic from z1z_{1} to z2z_{2}, and let pi=γ12Hip_{i}=\gamma_{12}\cap H_{i}. Let dijd_{ij} be the complex distance from WiinW_{i}^{in} to WjoutW_{j}^{out} along γij\gamma_{ij}. The spin frames WiinW_{i}^{in}, WioutW_{i}^{out} yield frames Fiin,FioutF_{i}^{in},F_{i}^{out} of unit parallel vector fields ViV_{i} on HiH_{i}.

Recall from Definition 4.3 that W2inW_{2}^{in} is obtained from W2outW_{2}^{out} by a rotation of π\pi about V2V_{2}, and W1outW_{1}^{out} is obtained from W1inW_{1}^{in} by a rotation of π-\pi about V1V_{1}. Define Y1outY_{1}^{out} to be the result of rotating W1inW_{1}^{in} by π\pi about V1V_{1}, so Y1outY_{1}^{out} and W1outW_{1}^{out} both project to F1outF_{1}^{out}, but differ by a 2π2\pi rotation.

The spin isometry which takes W1in(p1)W_{1}^{in}(p_{1}) to W2out(p2)W_{2}^{out}(p_{2}) also takes Y1out(p1)Y_{1}^{out}(p_{1}) to W2in(p2)W_{2}^{in}(p_{2}). Hence the complex distance d12d_{12} from W1in(p1)W_{1}^{in}(p_{1}) to W2out(p2)W_{2}^{out}(p_{2}) along γ12\gamma_{12} is equal to the complex distance from W2in(p2)W_{2}^{in}(p_{2}) to Y1out(p1)Y_{1}^{out}(p_{1}) along γ21\gamma_{21}. But since Y1outY_{1}^{out} and W1outW_{1}^{out} differ by a 2π2\pi rotation, this latter complex distance is d21+2πid_{21}+2\pi i. From d12=d21+2πid_{12}=d_{21}+2\pi i mod 4πi4\pi i we obtain λ12=λ21\lambda_{12}=-\lambda_{21}. ∎

If we have two spin frames adapted to to a common geodesic, and apply a homotopy MtPSL(2,)M_{t}\in PSL(2,\mathbb{C}) of isometries to them, for t[0,1]t\in[0,1], starting from the identity M0M_{0}, the complex distance between the spin frames remains constant; such a homotopy describes a point of the universal cover SL(2,)SL(2,\mathbb{C}). Hence complex distance between spin frames is invariant under the action of SL(2,)SL(2,\mathbb{C}). Similarly applying a homotopy to two spin-decorated horospheres and their common perpendicular geodesic, we observe that complex lambda length is also invariant under the action of SL(2,)SL(2,\mathbb{C}). In other words, if ASL(2,)A\in SL(2,\mathbb{C}) and (H1,W1),(H2,W2)(H_{1},W_{1}),(H_{2},W_{2}) are spin-decorated horospheres, then the complex lambda length from (H1,W1)(H_{1},W_{1}) to (H2,W2)(H_{2},W_{2}) is equal to the complex lambda length from A.(H1,W1)A.(H_{1},W_{1}) to A.(H2,W2)A.(H_{2},W_{2}).

We can now prove Theorem 2: given spinors κ1,κ22\kappa_{1},\kappa_{2}\in\mathbb{C}_{*}^{2}, and corresponding spin-decorated horospheres (Hi,Wi)=Φ2~Φ1~(κi)(H_{i},W_{i})=\widetilde{\Phi_{2}}\circ\widetilde{\Phi_{1}}(\kappa_{i}), the complex lambda length λ12\lambda_{12} form (H1,W1)(H_{1},W_{1}) to (H2,W2)(H_{2},W_{2}) satisfies

{κ1,κ2}=λ12.\{\kappa_{1},\kappa_{2}\}=\lambda_{12}.
Proof of Theorem 2.

Recalling that the spinor κ=(ξ,η)\kappa=(\xi,\eta) corresponds to a horosphere with centre ξ/η\xi/\eta, we observe that κ1,κ2\kappa_{1},\kappa_{2} are linearly dependent (over )\mathbb{C}) precisely when H1,H2H_{1},H_{2} have common centre. In other words, {κ1,κ2}=0\{\kappa_{1},\kappa_{2}\}=0 precisely when λ12=0\lambda_{12}=0. We can thus assume κ1,κ2\kappa_{1},\kappa_{2} are linearly independent.

First we prove the result when κ1=(1,0)\kappa_{1}=(1,0) and κ2=(0,1)\kappa_{2}=(0,1). From Proposition 3.9 then (H1,W1)(H_{1},W_{1}) is a spin lift of the decorated horosphere centred at \infty with height 11 and decoration specified by ii; and (H2,W2)(H_{2},W_{2}) is a spin lift of the decorated horosphere centred at 0 with Euclidean diameter 11 and decoration north-pole specified by ii. They are thus tangent at the point p=(0,0,1)p=(0,0,1), at which point W1inW_{1}^{in} and W2outW_{2}^{out} project to coincident frames, hence either coincide or differ by 2π2\pi.

To see that they coincide, we consider the following matrix ASL(2,)A\in SL(2,\mathbb{C}), regarded as the lift to the universal cover of the path MtPSL(2,)M_{t}\in PSL(2,\mathbb{C}) for t[0,π/2]t\in[0,\pi/2], starting at the identity:

A=[0110]SL(2,),Mt=±[costsintsintcost]PSL(2,).A=\begin{bmatrix}0&-1\\ 1&0\end{bmatrix}\in SL(2,\mathbb{C}),\quad M_{t}=\pm\begin{bmatrix}\cos t&-\sin t\\ \sin t&\cos t\end{bmatrix}\in PSL(2,\mathbb{C}).

Clearly A.κ1=κ2A.\kappa_{1}=\kappa_{2}, so by SL(2,)SL(2,\mathbb{C})-equivariance A.(H1,W1)=(H2,W2)A.(H_{1},W_{1})=(H_{2},W_{2}). Geometrically, in the upper half space model, MtM_{t} is a rotation of angle 2t2t about the oriented geodesic δ\delta from i-i to ii. Over t[0,π/2]t\in[0,\pi/2], the point pp and the vector ii specifying the decorations remain fixed, and the frame W1inW_{1}^{in} at pp rotates by π\pi about δ\delta to arrive at A.W1in=W2inA.W_{1}^{in}=W_{2}^{in}. Applying Definition 4.3, we then obtain the associated outward spin frame W2outW_{2}^{out} by a rotation of π-\pi about the decoration vector, i.e. about the same axis δ\delta. Thus indeed W1in(p)=W2out(p)W_{1}^{in}(p)=W_{2}^{out}(p), their complex distance is 0, and λ12=1\lambda_{12}=1.

Next we prove the result when κ1=(1,0)\kappa_{1}=(1,0) and κ2=(0,D)\kappa_{2}=(0,D) for some complex D0D\neq 0. In this case (H2,W2)(H_{2},W_{2}) is the spin lift of a decorated horosphere centred at 0, with Euclidean diameter |D|2|D|^{-2} and decoration north-pole specified by iD2iD^{-2}. The common perpendicular γ12\gamma_{12} runs from \infty to 0, intersecting H1H_{1} at p1=(0,0,1)p_{1}=(0,0,1) and H2H_{2} at p2=(0,0,|D|2)p_{2}=(0,0,|D|^{-2}). Thus the signed translation distance from p1p_{1} to p2p_{2} is 2log|D|2\log|D| and the rotation angle is given by argD2=2argD\arg D^{2}=2\arg D mod 2π2\pi; lifting to spin frames we show it is indeed 2argD2\arg D mod 4π4\pi. Consider again an ASL(2,)A\in SL(2,\mathbb{C}) lifting a path MtPSL(2,)M_{t}\in PSL(2,\mathbb{C}) from the identity,

A=[elog|D|iargD00elog|D|+iargD]=[D100D],Mt=±[et(log|D|+iargD)00et(log|D|+iargD)]A=\begin{bmatrix}e^{-\log|D|-i\arg D}&0\\ 0&e^{\log|D|+i\arg D}\end{bmatrix}=\begin{bmatrix}D^{-1}&0\\ 0&D\end{bmatrix},\quad M_{t}=\pm\begin{bmatrix}e^{-t(\log|D|+i\arg D)}&0\\ 0&e^{t(\log|D|+i\arg D)}\end{bmatrix}

where we take argD[0,2π)\arg D\in[0,2\pi) and t[0,1]t\in[0,1]. We have A.(0,1)=κ2A.(0,1)=\kappa_{2}, so AA sends Φ2~Φ1~(0,1)\widetilde{\Phi_{2}}\circ\widetilde{\Phi_{1}}(0,1) (i.e. (H2,W2)(H_{2},W_{2}) from the previous case) to (H2,W2)(H_{2},W_{2}) here. Geometrically MtM_{t} is a translation of length 2tlog|D|2t\log|D| and rotation of angle 2targD2t\arg D about γ12\gamma_{12}, so AA as a spin isometry translates by 2log|D|2\log|D| and rotates by 2argD2\arg D modulo 4π4\pi. Since the complex distance from W1inW_{1}^{in} to W2outW_{2}^{out} at p1p_{1} was zero in the previous case D=1D=1, the complex distance now becomes 2log|D|+2argDi2\log|D|+2\arg Di mod 4πi4\pi i. Thus λ12=D={κ1,κ2}\lambda_{12}=D=\{\kappa_{1},\kappa_{2}\}.

Finally, we prove the result for general linearly independent κ1,κ2\kappa_{1},\kappa_{2}. There exists ASL(2,)A\in SL(2,\mathbb{C}) such that A.κ1=(1,0)A.\kappa_{1}=(1,0) and A.κ2=(0,D)A.\kappa_{2}=(0,D), where D={κ1,κ2}D=\{\kappa_{1},\kappa_{2}\}. Applying this AA then the complex lambda length from (H1,W1)(H_{1},W_{1}) to (H2,W2)(H_{2},W_{2}) is equal to the complex lambda length from A.(H1,W1)A.(H_{1},W_{1}) to A.(H2,W2)A.(H_{2},W_{2}), which is {κ1,κ2}\{\kappa_{1},\kappa_{2}\} from the previous case. ∎

5 Hyperbolic geometry applications

The above theory can be applied to any situation involving horospheres in hyperbolic geometry, in up to 3 dimensions. Endowing each horosphere with a spin decoration, we obtain a spinor, and then applying the bilinear form {,}\{\cdot,\cdot\} gives us geometric information about horospheres.

As a first application we consider hyperbolic ideal tetrahedra, and prove the Ptolemy equation of Theorem 3. Take an ideal tetrahedron with vertices labelled 0,1,2,30,1,2,3, and a spin decoration (Hi,Wi)(H_{i},W_{i}) on each ideal vertex ii. We must show that the complex lambda lengths λij\lambda_{ij} from (Hi,Wi)(H_{i},W_{i}) to (Hj,Wj)(H_{j},W_{j}) satisfy

λ01λ23+λ03λ12=λ02λ13.\lambda_{01}\lambda_{23}+\lambda_{03}\lambda_{12}=\lambda_{02}\lambda_{13}. (5.1)
Proof of Theorem 3.

Let κi2\kappa_{i}\in\mathbb{C}^{2} be the spinor corresponding to (Hi,Wi)(H_{i},W_{i}). Let MM be the 2×42\times 4 complex matrices whose jj’th column is κj\kappa_{j}, and let MijM_{ij} be the 2×22\times 2 submatrix whose columns are κi\kappa_{i} and κj\kappa_{j} in order. Then detMij={κi,κj}=λij\det M_{ij}=\{\kappa_{i},\kappa_{j}\}=\lambda_{ij}, so the claimed equation becomes

detM01detM23+detM03detM12=detM02detM13,\det M_{01}\det M_{23}+\det M_{03}\det M_{12}=\det M_{02}\det M_{13},

which is a well known Plücker relation. ∎

Note that if we multiply any one of the spinors, say κi\kappa_{i} corresponding to (Hi,Wi)(H_{i},W_{i}), by a complex scalar cc, each term of the Ptolemy equation (5.1) involving index ii is also scaled by cc. For instance if we multiply κ1\kappa_{1} by cc then λ01,λ12,λ13\lambda_{01},\lambda_{12},\lambda_{13} are all multiplied by cc. In some sense then the choice of decorated horosphere at each vertex is a choice of gauge. The equation is in a certain sense, just the usual equation

z+z1=1,z+z^{\prime-1}=1, (5.2)

relating shape parameters for a hyperbolic ideal tetrahedron Δ\Delta, as we see next. By the shape parameter zez_{e} of Δ\Delta along an edge ee, we mean the complex number zz such that, if we place the two endpoints of ee at 0 and \infty, and place the remaining two ideal vertices at 11 and a point with positive imaginary part, then the final vertex lies at zz. By this definition, opposite edges of Δ\Delta have the same shape parameter, and the three pairs of shape parameters can be denoted z,z,z′′z,z^{\prime},z^{\prime\prime} such that z=11zz^{\prime}=\frac{1}{1-z} and z′′=z1zz^{\prime\prime}=\frac{z-1}{z}. In particular, (5.2) holds, and continues to hold if we cyclically permute (z,z,z′′)(z,z′′,z)(z,z^{\prime},z^{\prime\prime})\mapsto(z^{\prime},z^{\prime\prime},z).

Proposition 5.3.

Numbering the ideal vertices of Δ\Delta by 0,1,2,30,1,2,3 as in Figure 5, let the shape parameter of edge ijij by zijz_{ij}. Choose a spin-decorated horosphere (Hi,Wi)(H_{i},W_{i}) at ideal vertex ii and let λij\lambda_{ij} be the complex lambda length from (Hi,Wi)(H_{i},W_{i}) to (Hj,Wj)(H_{j},W_{j}). Then

z01=z23=λ02λ13λ03λ12,z02=z13=λ03λ12λ01λ23,z03=z12=λ01λ23λ02λ13.z_{01}=z_{23}=\frac{\lambda_{02}\lambda_{13}}{\lambda_{03}\lambda_{12}},\quad z_{02}=z_{13}=-\frac{\lambda_{03}\lambda_{12}}{\lambda_{01}\lambda_{23}},\quad z_{03}=z_{12}=\frac{\lambda_{01}\lambda_{23}}{\lambda_{02}\lambda_{13}}. (5.4)
Refer to caption
0
1
2
3
Figure 5: Tetrahedron with vertices labeled 0, 11, 22, 33.
Proof.

If we move a spin-decorated tetrahedron by a spin isometry, all shape parameters and complex lambda lengths remain invariant. Noting the orientation of Figure 5, we may place the ideal vertices 0,1,2,30,1,2,3 respectively at 0,,z,130,\infty,z,1\in\partial\mathbb{H}^{3} respectively, so z=z01z=z_{01}. With this arrangement then z=z01=z23z=z_{01}=z_{23}, z=z02=z13z^{\prime}=z_{02}=z_{13} and z′′=z03=z12z^{\prime\prime}=z_{03}=z_{12}. If we multiply a spinor κi\kappa_{i} corresponding to (Hi,Wi)(H_{i},W_{i}) by a complex scalar cc, the homogeneous expressions in lambda lengths in (5.4) are invariant. Thus it suffices to prove the claim for any single choice of spin decoration, or spinor, at each vertex. Take spinors κ0=(0,1)\kappa_{0}=(0,1), κ1=(1,0)\kappa_{1}=(1,0), κ2=(z,1)\kappa_{2}=(z,1), κ3=(1,1)\kappa_{3}=(1,1). By Theorem 2 then we calculate all complex lambda lengths as

λ01=1,λ02=z,λ03=1,λ12=1,λ13=1,λ23=z1.\lambda_{01}=-1,\quad\lambda_{02}=-z,\quad\lambda_{03}=-1,\quad\lambda_{12}=1,\quad\lambda_{13}=1,\quad\lambda_{23}=z-1.

This immediately gives the first equation. Permuting labels (0,1,2,3)(0,2,3,1)(0,1,2,3)\mapsto(0,2,3,1) on Δ\Delta, similarly permuting indices on each λ\lambda, and using the antisymmetry of λ\lambda, then gives the subsequent two equations. ∎

When the ideal tetrahedron Δ\Delta is degenerate and lies in a plane, by an isometry we may place Δ\Delta on the hyperbolic plane 2\mathbb{H}^{2} with 2={}\partial\mathbb{H}^{2}=\mathbb{R}\cup\{\infty\}, i.e. the upper half plane model inside the upper half space model. All vertices at infinity then lie in {}\mathbb{R}\cup\{\infty\}, and we may choose all spin directions to point perpendicular to 2\mathbb{H}^{2} in the following sense. Note that every horocycle HH centred at zz\in\mathbb{R}\cup\infty extends to a unique horosphere HH^{\prime} in 3\mathbb{H}^{3}, also centred at zz.

Definition 5.5.

Let HH be a horocycle in 23\mathbb{H}^{2}\subset\mathbb{H}^{3}. A planar spin decoration on HH is a spin decoration (Win,Wout)(W^{in},W^{out}) on HH^{\prime} such that Win,WoutW^{in},W^{out} project to frames specified by ii.

“Specified” here means north-pole specified, if HH^{\prime} is a sphere in the upper half space model.

Lemma 5.6.

A spinor yields a planar spin decoration on a horocycle if and only if it is real.

Proof.

The spin-decorated horosphere of κ=(ξ,η)\kappa=(\xi,\eta) will be a planar spin decoration on a horocycle if and only if the centre ξ/η{}\xi/\eta\in\mathbb{R}\cup\{\infty\} and the decoration direction i/η2i/\eta^{2} (if η0\eta\neq 0) or iξ2i\xi^{2} (if η=0\eta=0) is a positive real multiple of ii. This happens precisely when ξ,η\xi,\eta are real. ∎

Thus, we can reduce to two dimensions by considering real spinors, i.e. those in 2=2{(0,0)}\mathbb{R}_{*}^{2}=\mathbb{R}^{2}\setminus\{(0,0)\}. Then all complex distances dijd_{ij} between horospheres are real, so the λij=exp(dij/2)\lambda_{ij}=\exp(d_{ij}/2) are positive, giving Penner’s real lambda lengths between horocycles from [16, 17].

A horocycle in 2\mathbb{H}^{2} has two planar spin decorations, corresponding to the two spin lifts of frames specified by the ii direction. These two planar spin decorations correspond to two real spinors, which are negatives of each other.

6 Cluster algebra applications

We now develop the notions required to prove Theorem 4.

For reasons that will become apparent, we will consider 2\mathbb{H}^{2} via the upper half plane model, but to have the orientation induced by the normal vector in the ii direction in the upper half space model; this is the opposite to the usual orientation. Then the boundary circle 2{}\partial\mathbb{H}^{2}\cong\mathbb{R}\cup\{\infty\} obtains an orientation in the negative real direction.

Definition 6.1.

An ideal dd-gon is a collection of distinct points z1,,zdz_{1},\ldots,z_{d} in 2\partial\mathbb{H}^{2}, labelled in order around the oriented boundary 2\partial\mathbb{H}^{2}. A decoration on an ideal dd-gon is a choice of horocycle at each ziz_{i}.

We can join the points z1,,zdz_{1},\ldots,z_{d} (and back to z1z_{1}) successively by geodesics to form an ideal dd-gon in the usual sense, but we just need the ziz_{i}. Given the orientation on 2\partial\mathbb{H}^{2}, this means that the zi{}z_{i}\in\mathbb{R}\cup\{\infty\} satisfy either

zk>zk+1>>zd>z1>z2>>zk1z_{k}>z_{k+1}>\cdots>z_{d}>z_{1}>z_{2}>\cdots>z_{k-1} (6.2)

or

zk=andzk+1>>zd>z1>>zk1z_{k}=\infty\quad\text{and}\quad z_{k+1}>\cdots>z_{d}>z_{1}>\cdots>z_{k-1} (6.3)

for some kk.

In 3 dimensions, we lose the ordering on vertices of an ideal dd-gon, and instead use the following weaker notion. Again, our definition is just a sequence of ideal vertices, although we can imagine joining them by geodesics.

Definition 6.4.

A skew ideal dd-gon is a collection of distinct points z1,,zd3z_{1},\ldots,z_{d}\in\partial\mathbb{H}^{3}. A spin decoration on a skew ideal dd-gon is a choice of spin-decorated horosphere centred at each ziz_{i}.

We now use a special case of Penner’s definition in [16] in 2 dimensions, and generalise it naturally to 3 dimensions.

Definition 6.5.
  1. (i)

    The decorated Teichmüller space T~(d)\widetilde{T}(d) of ideal dd-gons is the space of all decorated ideal dd-gons, up to orientation-preserving isometries of 2\mathbb{H}^{2}.

  2. (ii)

    The decorated Teichmüller space T~3(d)\widetilde{T}^{3}(d) of skew ideal dd-gons is the space of all spin-decorated skew ideal dd-gons, up to spin isometries of 3\mathbb{H}^{3}.

In other words, the orientation-preserving isometry group PSL(2,)PSL(2,\mathbb{R}) acts on the space of decorated ideal dd-gons, and T~(d)\widetilde{T}(d) is the quotient. Similarly, the spin isometry group SL(2,)SL(2,\mathbb{C}) acts on the space of spin-decorated skew ideal dd-gons, and T~3(d)\widetilde{T}^{3}(d) is the quotient.

In the 2-dimensional case, with real spinors, the following statements demonstrate that a notion of total positivity has nice hyperbolic-geometric consequences. Similar ideas also appear in the physics literature, e.g. [1].

Definition 6.6.

A collection of spinors κ1,,κn\kappa_{1},\ldots,\kappa_{n} is totally positive if they are all real, and for all i<ji<j we have {κi,κj}>0\{\kappa_{i},\kappa_{j}\}>0.

Note that the totally positive condition implies that the κi\kappa_{i} are all linearly independent, so the corresponding horospheres are all centred at distinct points zi{}z_{i}\in\mathbb{R}\cup\{\infty\}; and by antisymmetry, when i>ji>j we have {κi,κj}<0\{\kappa_{i},\kappa_{j}\}<0.

Lemma 6.7.

Let d3d\geq 3. If κ1,,κd\kappa_{1},\ldots,\kappa_{d} are totally positive then the centres ziz_{i} of the corresponding horospheres form an ideal dd-gon. The planar spin-decorated horospheres of κ1,,κd\kappa_{1},\ldots,\kappa_{d} yield a map

{Totally positive d-tuples of spinors}{Decorated ideal d-gons}\left\{\text{Totally positive $d$-tuples of spinors}\right\}\longrightarrow\left\{\text{Decorated ideal $d$-gons}\right\}

which is surjective and 2-1, with the preimage of each ideal dd-gon of the form ±(κ1,,κd)\pm(\kappa_{1},\ldots,\kappa_{d}).

In other words, the totally positive condition forces the ziz_{i} to be in order around 2\partial\mathbb{H}^{2}, and we obtain a decorated ideal dd-gon. Conversely, any decorated ideal dd-gon is realised by precisely two dd-tuples of totally positive real spinors, which are negatives of each other.

Proof.

Letting κi=(ξi,ηi)\kappa_{i}=(\xi_{i},\eta_{i}) be totally positive we have

zizj=ξiηiξjηj={κi,κj}ηiηj.z_{i}-z_{j}=\frac{\xi_{i}}{\eta_{i}}-\frac{\xi_{j}}{\eta_{j}}=\frac{\{\kappa_{i},\kappa_{j}\}}{\eta_{i}\eta_{j}}. (6.8)

Supposing i<ji<j then, we have {κi,κj}>0\{\kappa_{i},\kappa_{j}\}>0, so ηi\eta_{i} and ηj\eta_{j} have the same sign when zi>zjz_{i}>z_{j}, and ηi\eta_{i} and ηj\eta_{j} have opposite signs when zi<zjz_{i}<z_{j}.

If z1,z2,z3z_{1},z_{2},z_{3} are real and satisfy z1<z2<z3z_{1}<z_{2}<z_{3}, or z2<z3<z1z_{2}<z_{3}<z_{1}, or z3<z1<z2z_{3}<z_{1}<z_{2}, then we obtain a contradiction. We show this in the case z1<z2<z3z_{1}<z_{2}<z_{3}; the other cases are similar. From z1<z2z_{1}<z_{2}, η1\eta_{1} and η2\eta_{2} have opposite signs. From z2<z3z_{2}<z_{3}, η2\eta_{2} and η3\eta_{3} have opposite signs. Thus η1\eta_{1} and η3\eta_{3} have the same sign. But η1\eta_{1} and η3\eta_{3} have opposite signs since z1<z3z_{1}<z_{3}, a contradiction.

This argument applies not just to z1,z2,z3z_{1},z_{2},z_{3} but to any zi,zj,zkz_{i},z_{j},z_{k} such that i<j<ki<j<k. These are precisely the cases in which zi,zj,zkz_{i},z_{j},z_{k} are not in order around 2\partial\mathbb{H}^{2}. Thus every triple of real numbers among the ziz_{i} is in order around 2\partial\mathbb{H}^{2}.

If all ziz_{i} are real, then considering the triples (z1,z2,z3)(z_{1},z_{2},z_{3}), (z1,z3,z4)(z_{1},z_{3},z_{4}), \ldots, (z1,zd1,zd)(z_{1},z_{d-1},z_{d}) shows that all ziz_{i} are in order around 2\partial\mathbb{H}^{2}, satisfying (6.2).

Suppose some zk=z_{k}=\infty, so zk=(ξk,0)z_{k}=(\xi_{k},0). We then have {κi,κk}=ηiξk\{\kappa_{i},\kappa_{k}\}=-\eta_{i}\xi_{k}. For i<ki<k then ηiξk\eta_{i}\xi_{k} is positive, so ηi\eta_{i} has the same sign as ξk\xi_{k}. Similarly, for i>ki>k, ηi\eta_{i} has opposite sign to ξk\xi_{k}. Thus if i<k<ji<k<j then ηi,ηj\eta_{i},\eta_{j} have opposite signs, so from (6.8) then zi<zjz_{i}<z_{j}. Applying the reasoning of the previous paragraph, it follows that (6.3) is satisfied.

Thus the ziz_{i} are in order around 2\partial\mathbb{H}^{2} and form an ideal dd-gon, and by Lemma 5.6 each κi\kappa_{i} yields a planar spin-decorated horosphere at ziz_{i}, hence a horocycle decoration in 2\mathbb{H}^{2}.

Conversely, suppose the ziz_{i} with horocycles HiH_{i} form a decorated ideal dd-gon in 2\mathbb{H}^{2}. Each ziz_{i} has two planar spin decorations, given by two real spinors of the form ±κi\pm\kappa_{i}. Choosing a sign for κ1\kappa_{1}, the requirement that each {κ1,κj}>0\{\kappa_{1},\kappa_{j}\}>0 forces a choice for each other κj\kappa_{j}. This yields two possible dd-tuples of real spinors describing the decorated ideal dd-gon; we will show they are both totally positive.

Suppose all ziz_{i} are real, satisfying (6.2) for some kk. Then by (6.8) ηi\eta_{i} has the same sign as η1\eta_{1} for 2ik12\leq i\leq k-1, and ηi\eta_{i} has a different sign from η1\eta_{1} for k+1idk+1\leq i\leq d. It follows that, for i<ji<j, ηi\eta_{i} and ηj\eta_{j} have the same sign precisely when zi<zjz_{i}<z_{j}, so by (6.2) {κi,κj}>0\{\kappa_{i},\kappa_{j}\}>0 for all i<ji<j, and the κi\kappa_{i} are totally positive. ∎

We can then give a description of T~(d)\widetilde{T}(d) in terms of totally positive spinors. The action of SL(2,)SL(2,\mathbb{R}) on real spinors extends to an action on dd-tuples as A.(κ1,,κd)=(A.κ1,,A.κd)A.(\kappa_{1},\ldots,\kappa_{d})=(A.\kappa_{1},\ldots,A.\kappa_{d}). We then obtain the following.

Proposition 6.9.

Let d3d\geq 3. The SL(2,)SL(2,\mathbb{R})-orbits of totally positive dd-tuples of real spinors are naturally bijective with T~(d)\widetilde{T}(d):

{Totally positive d-tuples of real spinors}SL(2,){Decorated ideal d-gons}PSL(2,)=T~(d).\frac{\left\{\text{Totally positive $d$-tuples of real spinors}\right\}}{SL(2,\mathbb{R})}\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\frac{\left\{\text{Decorated ideal $d$-gons}\right\}}{PSL(2,\mathbb{R})}=\widetilde{T}(d). (6.10)
Proof.

We first show the map of Lemma 6.7, sending totally positive dd-tuples to decorated ideal dd-gons, descends to a map as in (6.10). If two totally positive dd-tuples are related by the action of SL(2,)SL(2,\mathbb{R}), then by equivariance of the action of SL(2,)SL(2,)SL(2,\mathbb{R})\subset SL(2,\mathbb{C}), the resulting spin-decorated horospheres are related by a spin isometry of 23\mathbb{H}^{2}\subset\mathbb{H}^{3}, and dropping spin structures, the underlying decorated ideal dd-gons are related by an isometry in PSL(2,)PSL(2,\mathbb{R}).

Thus the map of (6.10) exists. It is also surjective since the map of Lemma 6.7 is. To see that it is injective, note the map of Lemma 6.7 is 2–1, with the two preimages of a given decorated dd-gon being negatives of each other. These two preimages are are related by the action of the negative identity in SL(2,)SL(2,\mathbb{R}), giving a unique preimage. ∎

We now define the Grassmannians we need. For background and context on positive Grassmannians, see e.g. [2, 13, 19, 21, 22]. Recall that the Grassmannian Gr𝔽(k,n)\operatorname{Gr}_{\mathbb{F}}(k,n) over a field 𝔽\mathbb{F} is the space of all kk-planes in 𝔽n\mathbb{F}^{n}. It can be realised as the quotient of Mat𝔽k(k,n)\operatorname{Mat}^{k}_{\mathbb{F}}(k,n), the space of all k×nk\times n matrices over 𝔽\mathbb{F} of rank kk, by the left action of GL(k,𝔽)GL(k,\mathbb{F}). A matrix represents the kk-plane spanned by its rows. The k×kk\times k minors of a matrix yield (nk)\binom{n}{k} projective coordinates on Gr𝔽(k,n)\operatorname{Gr}_{\mathbb{F}}(k,n) called Plücker coordinates. We only consider k=2k=2 and 𝔽=\mathbb{F}=\mathbb{R} or \mathbb{C}.

Definition 6.11.

Let Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d) denote the space of all 2×d2\times d real matrices with all Plücker coordinates positive. The positive Grassmannian Gr+(2,d)\operatorname{Gr}^{+}(2,d) is the quotient of Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d) by the left action of GL+(2,)GL^{+}(2,\mathbb{R}). The positive affine Grassmannian X+(d)X^{+}(d) is the affine cone on Gr+(2,d)\operatorname{Gr}^{+}(2,d).

The Plücker coordinates on a Grassmannian are only defined up to an overall factor, but they provide bona fide coordinates on the affine cone.

The affine cone on the Grassmannian Gr(2,d)\operatorname{Gr}_{\mathbb{R}}(2,d), as in [7, example 12.6], can be identified with the nonzero decomposable elements of Λ2(d)\Lambda^{2}(\mathbb{R}^{d}). The plane spanned by two rows R1,R2R_{1},R_{2} in a 2×d2\times d matrix is represented by R1R2R_{1}\wedge R_{2}, and the , and the action of AGL(2,)A\in GL(2,\mathbb{R}) is by

A.(R1R2)=(aR1+bR2)(cR1+dR2)=detAR1R2whereA=[abcd]A.(R_{1}\wedge R_{2})=(aR_{1}+bR_{2})\wedge(cR_{1}+dR_{2})=\det AR_{1}\wedge R_{2}\quad\text{where}\quad A=\begin{bmatrix}a&b\\ c&d\end{bmatrix}

Taking the quotient by GL(2,)GL(2,\mathbb{R}) thus identifies matrices whose corresponding decomposable elements of Λ2(d)\Lambda^{2}(\mathbb{R}^{d}) represent the same 22-plane in d\mathbb{R}^{d}. Taking the quotient by SL(2,)SL(2,\mathbb{R}) identifies matrices whose corresponding elements of Λ2(d)\Lambda^{2}(\mathbb{R}^{d}) are equal, and thus the affine cone on Gr(2,d)\operatorname{Gr}_{\mathbb{R}}(2,d) is the quotient of Mat2(2,d)\operatorname{Mat}^{2}_{\mathbb{R}}(2,d) by the left action of SL(2,)SL(2,\mathbb{R}). Restricting to matrices in Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d), taking the quotient by GL+(2,)GL^{+}(2,\mathbb{R}) again identifies matrices whose corresponding decomposable elements of Λ2(d)\Lambda^{2}(\mathbb{R}^{d}) which represent the same 22-plane, and taking the quotient by SL(2,)SL(2,\mathbb{R}) identifies matrices whose corresponding elements of Λ2(d)\Lambda^{2}(\mathbb{R}^{d}) are equal. Thus X+(d)X^{+}(d) is the quotient of Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d) by the left action of SL(2,)SL(2,\mathbb{R}).

Proof of Theorem 4(i).

We now have, by Proposition 6.9 and the above discussion

T~(n){Totally positive d-tuples of spinors}SL(2,)andX+(n)=Mat+(2,d)SL(2,).\widetilde{T}(n)\cong\frac{\{\text{Totally positive $d$-tuples of spinors}\}}{SL(2,\mathbb{R})}\quad\text{and}\quad X^{+}(n)=\frac{\operatorname{Mat}_{\mathbb{R}}^{+}(2,d)}{SL(2,\mathbb{R})}.

Placing a dd-tuple of real spinors (κ1,,κd)(\kappa_{1},\ldots,\kappa_{d}) as the columns of a 2×d2\times d matrix, the totally positive condition is that {κi,κj}>0\{\kappa_{i},\kappa_{j}\}>0 for i<ji<j. Each such {κi,κj}\{\kappa_{i},\kappa_{j}\} is then none other than the determinant of the 2×22\times 2 minor formed by columns ii and jj, i.e. the Plücker coordinate pijp_{ij}, so we precisely obtain the matrices in Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d). The actions of SL(2,)SL(2,\mathbb{R}) on totally positive dd-tuples and Mat+(2,d)\operatorname{Mat}_{\mathbb{R}}^{+}(2,d) are identical, so we obtain an identification T~(n)X+(n)\widetilde{T}(n)\longrightarrow X^{+}(n). By Theorem 2 each (complex) lambda length λij\lambda_{ij} on T~(n)\widetilde{T}(n) is equal to {κi,κj}\{\kappa_{i},\kappa_{j}\}, which we have seen is equal to the Plücker coordinate pijp_{ij} on X+(n)X^{+}(n). ∎

In a similar fashion over \mathbb{C}, we can consider the subvariety of the Grassmannian where all Plücker coordinates are nonzero.

Definition 6.12.

Let Mat(2,d)\operatorname{Mat}_{\mathbb{C}}^{*}(2,d) denote the space of all 2×d2\times d complex matrices with all Plücker coordinates nonzero. The nonzero Grassmannian Gr(2,d)\operatorname{Gr}_{\mathbb{C}}^{*}(2,d) is the quotient of Mat(2,d)\operatorname{Mat}_{\mathbb{C}}^{*}(2,d) by the left action of GL(2,)GL(2,\mathbb{C}). The nonzero affine Grassmannian X(d)X^{*}(d) is the affine cone on Gr(2,d)\operatorname{Gr}_{\mathbb{C}}^{*}(2,d).

Again the affine cone on Gr(2,d)\operatorname{Gr}_{\mathbb{C}}(2,d) can be identified with nonzero decomposable elements in Λ2(d)\Lambda^{2}(\mathbb{C}^{d}), and taking the quotient by SL(2,)SL(2,\mathbb{C}) identifies precisely those matrices whose corresponding elements of Λ2(d)\Lambda^{2}(\mathbb{C}^{d}) are equal. Thus X(d)X^{*}(d) is the quotient of Mat(2,d)\operatorname{Mat}_{\mathbb{C}}^{*}(2,d) by the left action of SL(2,)SL(2,\mathbb{C}).

Proof of Theorem 4(ii).

In a spin-decorated skew ideal dd-gon, at each ideal vertex ziz_{i} we have a spin-decorated horosphere corresponding to a spinor κi\kappa_{i}. The fact that all ziz_{i} are distinct (Definition 6.4) implies that for all iji\neq j we have {κi,κj}0\{\kappa_{i},\kappa_{j}\}\neq 0. By Definition 6.5, T~3(d)\widetilde{T}^{3}(d) is the space of all spin-decorated skew ideal dd-gons, up to spin isometries, so

T~3(d)={d-tuples of spinors with {κi,κj}0 for ij }SL(2,)andX(d)=Mat(2,d)SL(2,).\widetilde{T}^{3}(d)=\frac{\{\text{$d$-tuples of spinors with $\{\kappa_{i},\kappa_{j}\}\neq 0$ for $i\neq j$ }\}}{SL(2,\mathbb{C})}\quad\text{and}\quad X^{*}(d)=\frac{\operatorname{Mat}_{\mathbb{C}}^{*}(2,d)}{SL(2,\mathbb{C})}.

Again, putting the dd spinors as the columns of a matrix and noting that the SL(2,)SL(2,\mathbb{C}) actions are identical gives an identification T~3(d)X(d)\widetilde{T}^{3}(d)\longrightarrow X^{*}(d), and each complex lambda length λij={κi,κj}\lambda_{ij}=\{\kappa_{i},\kappa_{j}\} is equal to the corresponding Plücker coordinate pijp_{ij}. ∎

References

  • [1] N. Arkani-Hamed, J. Bourjaily, F. Cachazo, A. Goncharov, A. Postnikov, and J. Trnka, Grassmannian geometry of scattering amplitudes, Cambridge, 2016.
  • [2] K. Baur, Grassmannians and cluster structures, Bull. Iranian Math. Soc. 47 (2021), S5–S33.
  • [3] F. Chapoton, S. Fomin, and A. Zelevinsky, Polytopal realizations of generalized associahedra, vol. 45, 2002, Dedicated to Robert V. Moody, pp. 537–566.
  • [4] V. Fock and A. Goncharov, Moduli spaces of local systems and higher Teichmüller theory, Pub. Math. IHES (2006), 1–211.
  • [5] S. Fomin, M. Shapiro, and D. Thurston, Cluster algebras and triangulated surfaces. I. Cluster complexes, Acta Math. 201 (2008), no. 1, 83–146.
  • [6] S. Fomin and D. Thurston, Cluster algebras and triangulated surfaces Part II: Lambda lengths, Mem. AMS 255 (2018), no. 1223.
  • [7] S. Fomin and A. Zelevinsky, Cluster algebras. II. Finite type classification, Invent. Math. 154 (2003), no. 1, 63–121.
  • [8]  , YY-systems and generalized associahedra, Ann. of Math. (2) 158 (2003), no. 3, 977–1018.
  • [9] L. R. Ford, Fractions, Amer. Math. Monthly 45 (1938), no. 9, 586–601.
  • [10] S. Garoufalidis, D. Thurston, and C. Zickert, The complex volume of SL(n,){\rm SL}(n,\mathbb{C})-representations of 3-manifolds, Duke Math. J. 164 (2015), no. 11, 2099–2160.
  • [11] C. Geiss, B. Leclerc, and J. Schröer, Partial flag varieties and preprojective algebras, Ann. Inst. Fourier (Grenoble) 58 (2008), no. 3, 825–876.
  • [12] M. Gekhtman, M. Shapiro, and A. Vainshtein, Cluster algebras and Weil-Petersson forms, Duke Math. J. 127 (2005), no. 2, 291–311.
  • [13] G. Lusztig, Total positivity in reductive groups, Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 531–568.
  • [14] D. V. Mathews and J. Purcell, Ptolemy equations for hyperbolic 3-manifolds, in preparation.
  • [15] D. V. Mathews and O. Zymaris, Spinors and Descartes’ theorem, in preparation.
  • [16] R. C. Penner, The decorated Teichmüller space of punctured surfaces, Comm. Math. Phys. 113 (1987), no. 2, 299–339.
  • [17]  , Decorated Teichmüller theory, QGM Master Class Series, European Mathematical Society (EMS), Zürich, 2012, With a foreword by Yuri I. Manin.
  • [18] R. Penrose and W. Rindler, Spinors and space-time. Vol. 1, Cambridge, 1984.
  • [19] A. Postnikov, Total positivity, Grassmannians, and networks, arxiv:0609764.
  • [20] J. Scott, Grassmannians and cluster algebras, Proc. London Math. Soc. (3) 92 (2006), no. 2, 345–380.
  • [21] L. K. Williams, The positive Grassmannian, the amplituhedron, and cluster algebras, arxiv:2110.10856.
  • [22]  , Cluster algebras: an introduction, Bull. Amer. Math. Soc. (N.S.) 51 (2014), no. 1, 1–26.
  • [23] C. Zickert, Ptolemy coordinates, Dehn invariant and the AA-polynomial, Math. Z. 283 (2016), no. 1-2, 515–537.