Spinors and horospheres
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 to vectors , 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 on the future light cone , together with a 2-plane tangent to containing , 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 in Minkowski space of one lower dimension , to horocycles in the hyperbolic plane , 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).
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, -equivariant correspondence between nonzero spinors, and spin-decorated horospheres in hyperbolic 3-space .
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 are isometric to the Euclidean plane. A spin decoration is, roughly, a “spin lift” of such a decoration, where rotating the direction by is not the identity, but rotation by 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 . Regarding the sphere at infinity of as in the usual way, the horosphere corresponding to has centre at . If then is a horizontal plane in at height , and its decoration points in the direction . If then is a Euclidean sphere in of Euclidean diameter , and its decoration at its highest point (“north pole”) points in the direction . See Figure 2 (right). We prove this in Proposition 3.9.
As indicated, both sides of this correspondence have -actions. The action on is by the standard action of matrices on vectors. The action on horospheres is via the action of as orientation-preserving isometries of ; this action lifts to to preserve spin structures.
Penrose and Rindler in [18] define a complex-valued antisymmetric bilinear form on spinors, given by the determinant; it can also be regarded as the standard complex symplectic form on . On the other hand, given two horospheres , we may measure the signed distance from one to the other along their mutually perpendicular geodesic. As detailed in Section 4, we may also measure an angle between spin decorations, which is well defined modulo . We regard as a complex distance between the two horospheres. See Figure 3.
The correspondence of Theorem 1 extends further to relate these structures, as follows.
Theorem 2.
Given two spinors , suppose the corresponding spin-decorated horospheres have complex distance . Then
In the 2-dimensional context, the distance between horocycles is just a real number , and the quantity is known as a lambda length [16]. Thus, the standard bilinear form on spinors computes lambda lengths between corresponding spin-decorated horospheres.
Example.
Take and . In the upper half space model, corresponds to a horosphere centred at , so appears as a horizontal plane, at height and with decoration in the direction . Similarly, the horosphere of has centre , appearing as a sphere centred at , with Euclidean diameter , and direction at its highest point in the direction . These two horospheres are tangent at the point , where their decorations align. It turns out their spin directions also align, so and hence .
Multiplying by a complex number , with and real, its horosphere remains centred at , but the horizontal plane moves to height , translated upwards from its original position by hyperbolic distance , and its decoration is rotated by . From to we then have and , so .
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 , following Penrose–Rindler, we obtain points on the positive light cone in via
(1.1) |
which uses the correspondence between Hermitian matrices and Minkowski space given by Pauli matrices. The flag of has flagpole along the ray of determined by (1.1), and 2-plane defined by the derivative of this map in a certain direction depending on ; we show this is a variation on the Penrose–Rindler construction. The correspondence between spinors and flags is -equivariant, where acts on flags via its action on in standard fashion as . In Section 2 we describe these constructions precisely, along with explicit calculations, and details of -equivariance, which we have not seen proved elsewhere in the literature.
From a point on the figure light cone , Penner’s construction in [16] is to consider the affine plane in Minkowski space consisting of all satisfying
where is the standard Lorentzian metric. This affine plane intersects the hyperboloid model of hyperbolic space in a horosphere . This construction works in any dimension, and in , we obtain an affine 3-plane which intersects the hyperboloid model of in a 2-dimensional horosphere. We show that when arises from a spinor , this affine 3-plane contains an affine 2-plane parallel to the flag 2-plane of , and this affine 2-plane intersects the horosphere in an parallel oriented line field on , yielding the picture of Figure 2 (left). Again, all constructions are -equivariant. Precise details are given in Section 3.
Finally, all these constructions lift, in appropriate sense, to spin double covers, and remain -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 . Number the vertices . We consider a spin decoration on the tetrahedron, consisting of a spin-decorated horosphere at each ideal vertex . There is then a complex lambda length from to . Each 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 in a spin-decorated ideal tetrahedron satisfy
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 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 .
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 , which can be arranged into a matrix. The Ptolemy equation is then just the Plücker relation between determinants of a matrix. Alternatively, it can be seen as the relation
satisfied by the spinor , 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 and are both integers then the horocycles obtained in the upper half plane model of 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 can be arranged into a matrix. Considering those 4 spinors up to the common action of corresponds to considering such a up to isometry. And considering appropriate matrices up to a left action by 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 in terms of Grassmannians. Clusters in this case are in bijection with triangulations of an -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 , the affine cone over the Grassmannian in particular, the cluster algebra there denoted in type is isomorphic to the -form of the coordinate ring , 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 introduced by Penner [16], using lambda lengths. In the particular case of an -gon, the lambda lengths of the diagonals provide cluster variables for the cluster algebra of type [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 over the Grassmannian , and from the decorated Teichmüler space . 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 . The correspondence of Theorem 1 yields the following identifications.
-
(i)
The decorated Teichmüller space of ideal -gons is identified with the affine cone on the positive Grassmannian .
-
(ii)
The decorated Teichmüller space of ideal skew -gons in is identified with the affine cone on the subvariety of the complex Grassmannian 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 spinors describes the ideal vertices of an ideal -gon (in 2 dimensions), or an ideal skew -gon (in 3 dimensions); but on the other hand, we may place the as the columns of a matrix. The appropriate decorated Teichmüller space is then given by the certain (spin) isometry classes of such ideal (skew) -gons, which is an orbit space of a -tuple of spinors . The corresponding set of orbits of matrices gives the affine cone on the appropriate Grassmannian.
It is not difficult to vary the conditions on -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 . 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 , which we regard as a complex symplectic vector space, with complex symplectic form denoted
following [18]. We denote spinors by or similar. Given and then
We write for the above determinant. We denote by the set of nonzero spinors.
For the purposes of linear algebra, we regard as a column vector and write for the corresponding row vector. The adjoint is then a row vector.
We map spinors into the set of Hermitian matrices, or equivalently into Minkowski space . We take to have coordinates and metric , denoted . We observe and are isomorphic 4-dimensional real vector spaces and we identify them in a standard way (perhaps the constant is slightly unorthodox)
The right hand expression is , where the are the Pauli matrices. If a point corresponds to then we observe and . The light cone corresponds to with determinant zero, and the future light cone corresponds to satisfying and . We define the celestial sphere to be the intersection of with the 3-plane .
Definition 2.1 ([18]).
The map from to is defined by .
In other words,
We observe that the image of has determinant zero, and its diagonal entries are , so that its trace is non-negative. Indeed it is not difficult to show . Thus maps a 4-(real)-dimensional domain onto a 3-dimensional image. The fibres are circles; it is not difficult to show that iff for some real . Indeed, on each 3-sphere in given by with fixed at some constant , restricts to the Hopf fibration onto the 2-sphere in given by . Thus is the cone on the Hopf fibration.
In order not to lose information, we extend to a map including tangent data. Given a tangent vector in the real tangent space , we write for the derivative of at in the direction . Since, for real ,
we have
(2.2) |
In the abstract index notation of [18], this directional derivative is . At each point we will build a flag structure using the derivative in a certain direction .
Definition 2.3.
The function is given by . In other words,
Let us attempt to motivate this definition. Penrose–Rindler use a spinor forming a spin frame, or standard symplectic basis, with , i.e. so that . They then form a 2-plane defined by the bivector where and . These two vectors are our and . But the same oriented 2-plane is obtained using any positive multiple of such , so we could equally fix simply to be positive real. Choosing to make negative real, or positive/negative imaginary, works also for our purposes. Our choice of ensures is negative imaginary. Though somewhat arbitrary, this works well for our purposes.
Another perspective on is obtained by identifying with the quaternion . Then . On the centred at the origin in through , the tangent space at has basis . In the direction lies the fibre , and is constant; is another tangent vector to this .
We now define the type of flag structure we need.
Definition 2.5.
An oriented flag of signature in a real vector space is an increasing sequence of subspaces
where , and for , the quotient is endowed with an orientation.
Definition 2.6.
A pointed oriented null flag, or just flag, consists of a point and an oriented flag in of signature , such that
-
(i)
and the orientation on is towards the future (i.e. from towards ),
-
(ii)
is a tangent plane to .
The set of flags is denoted .
Thus is a on flagpole , which runs towards the future along the light cone; and the flag plane is a tangent plane to the light cone, with its relative orientation equivalent to choosing the half-plane to one side of or the other. Note that contains no timelike vectors, and generates the unique 1-dimensional lightlike subspace of . The tangent space to at is defined by the equation , i.e. is the (Minkowski-)orthogonal complement . Thus .
Given linearly independent and , we denote by the flag given by , the line oriented from the origin towards , the plane spanned by and , and the orientation on induced by . We observe that two flags so given , are equal if and only if and there exist real such that , where (which are necessarily nonzero) have opposite sign.
Note that is diffeomorphic to , where is the unit tangent bundle of : a point of describes a future-oriented ray in , a unit tangent vector there describes a relatively oriented 2-plane, and the factor fixes along the ray. Since we also have
Our version of Penrose–Rindler null flags can now be defined as the following map, upgrading .
Definition 2.7.
The map maps nonzero spinors to (pointed oriented null) flags via
Thus the point yields the flagpole, and the derivative of in the direction yields the relatively oriented flag plane. We verify that is (real-)linearly independent from using (2.4): if for some real then ; both sides of this equation being the product of a and matrix, the corresponding matrices must be proportional, say for some real ; in components then and , so and , so that , a contradiction.
Lemma 2.8.
For two spinors , the following are equivalent:
-
(i)
is negative imaginary (just like );
-
(ii)
where is complex and is real positive;
-
(iii)
Proof.
If is negative imaginary then for some positive , and any two vectors yielding the same value for differ by a complex multiple of . This shows (i) implies (ii), and the converse is clear.
If then by linearity of the derivative, . The derivative of in the direction is proportional to , and the derivative in the direction is zero (pointing along a fibre of ). Thus the derivatives in the and directions span the same plane when taken together with ; indeed, as , the same relatively oriented plane. In fact, this condition is equivalent to spanning the same relatively oriented plane. ∎
The spaces , and all have natural actions; in all cases we denote the action of by a dot. An acts on by the defining representation, , yielding a symplectomorphism:
since . The same acts on by , which in yields in the standard way the linear maps of , i.e. those which preserve the Minkowski metric and space and time orientation. The action on induces orientation-preserving actions on and planes in , yielding an action on , so that . Essentially by definition is equivariant with respect to these actions,
and we have an equivariance property on its derivatives
since . We now show the equivariance property extends to ; we have not seen a proof of this in the existing literature.
Lemma 2.9.
The map is equivariant with respect to the actions on and .
Proof.
We have so
by equivariance of and its derivative. Now as is symplectic, , which is negative imaginary, so by Lemma 2.8 then . ∎
It is possible to express explicitly the linear dependence implied by the equality of the flags and : a direct computation verifies the (perhaps surprising) identity
We can compute completely explicitly.
Lemma 2.10.
Let . Then in we have
and
The fact that has zero -coordinate follows from being tangent to the centred at the origin through , which maps under to the given by the intersection of with a plane at constant .
We denote by the hyperboloid model of hyperbolic 3-space
and by the boundary at infinity of . So and is naturally bijective with the celestial sphere .
Indeed, projectivising yields an the boundary at infinity and under this projectivisation, 2-planes tangent to containing a ray of correspond bijectively with tangent lines at points of . Moreover, relatively oriented planes containing a ray of correspond bijectively with tangent directions at points of .
The orientation-preserving isometry group of acts transitively on the future light cone , and indeed acts transitively on the tangent directions at points of . Further, if we take an oriented flag consisting of a future-oriented line of and a relatively oriented 2-plane tangent to , then there is an element of fixing (and its orientation) and (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 . In other words, acts transitively on , and the action factors through .
Taking , we have , which we denote , and by Lemma 2.10, is the flag with basepoint and 2-plane spanned by and . Thus as we multiply by to move through a fibre of , the flag rotates about a fixed pointed flagpole twice as fast. It follows that takes the value of each such flag exactly twice.
Using the equivariance of and the transitive action of , the same applies for the flags based at any point on . It follows that is smooth, surjective and 2–1. Moreover, the stabiliser of a flag in is trivial, so that acts freely and transitively on . Topologically is a map which is a double cover.
3 From Minkowski space to horospheres
We have now built the maps and in the commutative diagram
where the downwards arrow forgets all structure of a flag except its point on . In this section we define the maps , and the spaces , which involve horospheres and decorations.
Horospheres in the hyperboloid model are given by the intersection of with certain affine 3-planes in . Any affine 3-plane in is given by satisfying an equation of the form , where is a (Minkowski-)normal vector to the plane and is a real constant. We call such an affine 3-plane lightlike if its normal is lightlike. We observe that a lightlike 3-plane can be defined by an equation where ; if then this plane intersects in a horosphere, and if the plane is disjoint from . Normalising such equations by requiring the constant to be , i.e. , then gives a bijection between points and horospheres. We denote the set of horospheres in by .
Definition 3.1 ([16]).
The map sends to the horosphere defined by . The map sends to the point at infinity of .
Thus the map is a bijection. Indeed, it is a diffeomorphism: , with an -family of horospheres at each point at infinity in . Any horosphere has a unique point at infinity in , which we also call its centre. The map can be regarded as the projectivisation map or projection to the celestial sphere .
Note acts naturally on (as on and ) in the standard way, via linear maps of , and hence also on and , and we again denote all actions via a dot. We observe an sends the horosphere , defined by , to the horosphere defined by . Since the action of preserves the Minkowski metric, this horosphere is also given by . In other words, so that is -equivariant. Forgetting the horospheres and recording only points at infinity, similarly is -equivariant.
We now consider the intersection of a horosphere with a flag. So consider a horosphere for some , and consider a flag based at the same , given by the oriented sequence . The horosphere is the intersection of , given by , with the plane ; hence at a point , its tangent space is given by . The intersection of the horosphere with the flag plane at will thus be given by
since . Now the intersection is the intersection of a spacelike 3-plane , and the 2-plane , so it is either 1- or 2-dimensional. But if it were 2-dimensional then we would have ; but contains a timelike vector , while is spacelike. Thus the intersection is 1-dimensional and spacelike.
Moreover, the orientation on is an orientation on , and thus any vector in not in obtains an orientation, depending on the side of to which it lies. The intersection is spacelike, hence not equal to . Thus we may regard the intersection of the horosphere with the flag plane as defining an oriented line tangent to the horosphere at each point. In other words, we obtain an oriented line field on . We denote by the set of horospheres with oriented line fields.
Definition 3.2.
The map sends a flag to the horosphere , with the oriented line field defined at each point by .
An acts on : linear maps in are orientation-preserving isometries of , sending horospheres to horospheres, with their derivatives sending oriented line fields to oriented line fields. Since the -actions on and are both via linear maps of acting on , is -equivariant.
It is well known that a horosphere is isometric to a Euclidean 2-plane. The parabolic orientation-preserving isometries of fixing 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 is parallel if it is invariant under Euclidean translations (i.e. under the action of all parabolic isometries fixing ).
A decorated horosphere is a horosphere with a parallel oriented line field. The set of all decorated horospheres is denoted .
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 for a simple but useful example.
Lemma 3.4.
is the horosphere in which has point at infinity in the direction along , passing through , with the oriented parallel line field pointing in the direction at .
Proof.
We have , so that is the horosphere given by , which is indeed the horosphere . From Lemma 2.10 the flag is given by , so the flag 2-plane is spanned by and , with relative orientation on given by .
Now the parabolic subgroup
(3.5) |
fixes and acts simply transitively on . Denoting by the matrix in with upper right entry , the points of are parametrised by ; letting we have . We calculate the action of on to be
where
Thus we calculate
and moreover for we have . At the line field of is given by . Now is the 3-plane given by equation
(3.6) |
while is spanned by and , hence defined by and . Thus is defined by , and , hence spanned by . Since the orientation on is given by , the oriented line field of at is directed by . In particular, the oriented line field of at is directed by .
Now, if we apply to the vector directing the line field at a point of , we obtain the vector at . Thus the oriented line field is parallel. ∎
In fact in the above calculation we observe that and . This shows explicitly that the parabolic subgroup preserves the flag plane , and in fact acts as the identity on both and .
In fact this example is generic enough to give the following.
Lemma 3.7.
The map is a diffeomorphism .
In other words, the oriented line field of any flag is parallel, and provides a smooth correspondence between flags and decorated horospheres.
Proof.
First we show always yields parallel oriented line fields. Lemma 3.4 shows this when is applied to . But the action of is transitive on , and the action of on (hence on horospheres) is by isometries, and is -equivariant. Any flag in is thus of the form for some , so , which has parallel oriented line field.
Next we show that sends the flags of the form bijectively to decorations on . These flags are those of the form
as calculated above. Denote the latter vector by , so . Then is the horosphere , with oriented parallel line field at given by the intersection of the flag 2-plane with . Since is given by (equation ((3.6)) with ), which contains , the oriented line field of at is directed by . As increases say from to , both the flag through and the decoration on rotate through a full , with providing a bijection.
We have already seen that provides a bijection between and ; using the transitivity of on and , and equivariance of , it follows that provides a bijection between the flags based at each , and decorations on the corresponding horospheres .
Thus is a bijection. It and its inverse are clearly smooth, once and are given their natural smooth structures. We have already seen . The space of horospheres is naturally , and decorations can be given by unit tangent vectors to the sphere at infinity, so that . ∎
We now consider our horospheres in the upper half space model of , given in the usual way as
As usual we identify the plane with and with , and coordinates with . In , horospheres centred at appear as horizontal planes; we call the -coordinate of this plane the height of the horosphere. Horospheres centred at other points appear as spheres tangent to ; we call the maximum of on the sphere the Euclidean diameter of the horosphere.
We proceed from to via the disc model . We have the standard isometries given by
(3.8) |
where in the latter map we regard as the standard and as . Of course -actions carry through equivariantly to each model as isometries, and on the action is via Möbius transformations in the usual way,
We now introduce some terminology to describe decorations, i.e. parallel oriented line fields, on horospheres in . A horosphere centred at is a horizontal plane parallel to , 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 -coordinate (its “north pole”); since the tangent plane there is also parallel to , 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 maps under to a decorated horosphere whose centre is at in the upper half space model.
-
(i)
If then the horosphere has Euclidean diameter , and decoration north-pole specified by ,
-
(ii)
If , then the horosphere has height , and decoration specified by .
In particular, forgetting the decorations, the above proposition gives an explicit description of . And forgetting all but the centres of the horospheres, it yields .
Proof.
Letting , , is given in Lemma 2.10. . Then , for the hyperboloid model, just projectivises the rays of to points; taking for the point on each ray with gives on as
The centre of the horosphere on is then, using (3.8),
From Lemma 3.4, , in the hyperboloid model, is the horosphere centred at , passing through , and at has decoration in the direction . In , this corresponds to the horosphere centred at , passing through , and having decoration in the direction there. In , this corresponds to the horosphere centred at , passing through , and having decoration in the direction at that point. In other words, it has height and decoration specified by .
The decorated horospheres can now be found in general using -equivariance. Observe that
(3.10) |
Thus the decorated horosphere of is obtained from the decorated horosphere of by applying the Möbius transformation ; hence it is centred at , has Euclidean diameter , and is north-pole specified by . Similarly, the decorated horosphere of , for , is obtained from that of by applying , hence is centred at , has height , and is specified by . And the decorated horosphere of , for , is obtained from that of by applying , hence is centred at , has Euclidean diameter , and is north-pole specified by . ∎
Thus, if we multiply a spinor by a complex number , with and real, the effect on the corresponding horosphere is to translate it by distance along any geodesic perpendicular to oriented towards its centre, and rotate the decoration by about .
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 refers to hyperbolic 3-space, regardless of model.
We use the cross product in in the elementary sense that if are tangent to at a common point , making an angle of , then is tangent to at , has length , and points in the direction perpendicular to and given by the right-hand rule.
A horosphere in (like any oriented surface in a 3-manifold) has two normal directions: we call the direction towards its centre outward (“pointing out of ”), and the direction away from its centre inward (“pointing into ”). There are well-defined outward and inward unit normal vector fields along , which we denote respectively.
By a frame we mean a right-handed orthonormal frame at a point in , i.e. a triple of orthogonal unit vectors such that . The collection of frames then forms a principal -bundle over which we denote
We may take its spin double (universal) cover, which we denote
which is a principal -bundle. We refer to points of as spin frames. Each point in has two lifts in , i.e. each frame lifts to two spin frames.
The group of orientation-preserving symmetries of is naturally isomorphic to , and acts simply transitively on . Choosing a basepoint in then we may obtain an explicit identification , given by for .
Similarly, acts simply transitively on . And the identification lifts to double covers, after we choose a lifted basepoint , giving an explicit diffeomorphism as . The two matrices lifting then correspond to the two spin frames , lifting the frame . These two spin frames are related by a rotation about any axis at their common point. We can regard elements of as spin isometries; each isometry in lifts to two spin isometries, which differ by a rotation. Since is the universal cover of the isometry group , we can also regard elements of as homotopy classes of paths of isometries starting at the identity.
From a decoration on a horosphere , normalised to a parallel unit tangent vector field on , we can then construct frame fields along as follows.
Definition 4.1.
Let be a unit parallel tangent vector field on a horosphere .
-
(i)
The inward frame field of is the frame field on given by .
-
(ii)
The outward frame field of is the frame field on given by .
Indeed a decorated horosphere is uniquely specified by its inward and outward frame fields and so we can denote a decorated horosphere by where is the pair of frames .
A frame field is a continuous section of along , and it has two lifts to .
Definition 4.2.
An outward (resp. inward) spin decoration on is a continuous lift of an outward (resp. inward) frame field from to .
From the inward frame field of a unit parallel vector field on , one can rotate the frame at each point of by an angle of or about to obtain the outward frame field of , 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 about , which will result in an outward spin decoration. However, rotations of or about yield distinct results, related by a 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.
-
(i)
Let be an outward spin decoration on . The associated inward spin decoration is the spin decoration obtained by rotating by angle about at each point of .
-
(ii)
Let be an inward spin decoration on . The associated outward spin decoration is the spin decoration obtained by rotating by angle about at each point of .
We observe that associated spin decorations come in pairs , each associated to the other.
Definition 4.4.
A spin decoration on a horosphere is a pair of associated inward and outward spin decorations. We denote a spin-decorated horosphere by , and denote the set of spin-decorated horospheres by .
Note that under the identification , with an appropriate choice of basepoint frame, the parabolic subgroup of equation (3.5) (or more precisely its image in ) corresponds to all the frames of the outward frame field of . The cosets of then correspond bijectively with decorated horospheres. Similarly, under the identification with an appropriate choice of basepoint, the cosets of correspond bijectively with spin-decorated horospheres:
We now consider lifts of the maps
Topologically, we have , we have seen , and we have seen is a double cover and is a diffeomorphism. Indeed, all the spaces here are bundles over and the maps are bundle maps, which in an appropriate sense are the identity on the base space . The spaces and both have fundamental group , and we can consider their double (hence universal) covers. A nontrivial loop in is given by fixing a flagpole and rotating a flag through ; in the double cover, rotating the flag through is no longer a loop, but rotating the flag through gives a loop.
Definition 4.5.
The double cover of the space of flags is denoted . We call its elements spin flags.
Our spin flags are the null flags of [18].
A nontrivial loop in is given by fixing a horosphere and rotating its decoration through . In the double cover, a rotation through is not a loop but a rotation through gives a loop. In other words, the double cover of is . Choosing basepoints (arbitrarily) one then obtains lifts such that the diagram
commutes, where the downwards arrows are double covering maps. The action of (which is simply connected) lifts to actions on these covers and all maps remain -equivariant.
Thus a spinor maps under to a spin-decorated horosphere lifting the decorated horosphere described in Proposition 3.9. Multiplying by , with and real, still translates it towards its centre and rotates the decoration by , but now the rotation is taken modulo .
We can now prove Theorem 1, that there is an explicit smooth bijective correspondence between and .
Proof of Theorem 1.
At the end of Section 2 we observed that is a smooth double cover, topologically . In Lemma 3.7 we showed is a diffeomorphism. Their lifts and are then both diffeomorphisms, topologically . We have defined these maps explicitly. We have also shown all maps are -equivariant. Thus 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 be a point on an oriented geodesic in . A frame at is adapted to if is positively tangent to . A spin frame at is adapted to if it is the lift of a frame adapted to .
Now if we have two points on an oriented geodesic , and frames at each , adapted to , then there is then a screw motion along which takes to as follows. Being adapted to , the first vectors and in each frame point along . Parallel translation along from to takes to a frame at which agrees with in its first vector. This translation is by a signed distance which we regard as positive or negative according to the orientation on . A further rotation of some angle about (signed using the orientation of ) then moves to . Note that is only well defined modulo . However we may repeat this process with spin frames, and then is well defined modulo .
Definition 4.7.
Let be frames, or spin frames, at points on an oriented geodesic , adapted to . The complex distance from to is , where a translation along of signed distance , followed by a rotation about of angle , takes to .
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 ; between spin frames, it is well defined modulo .
We can now define complex lambda lengths between decorated and spin-decorated horospheres. Let be horospheres, let be the centre of , and let be the oriented geodesic from to . Thus and are the two orientations of the unique common perpendicular to the horospheres. Let . If the are decorated, we have pairs of inward and outward frame fields on each , and note that and are both adapted to . If the are spin-decorated, we have pairs of associated inward and outward spin decorations on each , and we note that and are adapted to .
Definition 4.8.
-
(i)
If and are decorated horospheres, the complex lambda length from to is
where is the complex distance from to .
-
(ii)
If and are spin-decorated horospheres, the complex lambda length from to is
where is the complex distance from to .
When the horospheres and have a common centre, then the complex lambda length between them is zero in either case.
Note that for decorated horospheres, is only well defined modulo , so is only well defined up to sign. For spin-decorated horospheres however is well defined modulo , so is a well defined complex number, and indeed we have a well defined function .
We observe that 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 and hence .
In fact, as we now see, is antisymmetric.
Lemma 4.9.
Let , be spin-decorated horospheres, and let be the complex lambda length from to . Then .
Proof.
If have common centre . So we may assume have distinct centres . As above, let be the oriented geodesic from to , and let . Let be the complex distance from to along . The spin frames , yield frames of unit parallel vector fields on .
Recall from Definition 4.3 that is obtained from by a rotation of about , and is obtained from by a rotation of about . Define to be the result of rotating by about , so and both project to , but differ by a rotation.
The spin isometry which takes to also takes to . Hence the complex distance from to along is equal to the complex distance from to along . But since and differ by a rotation, this latter complex distance is . From mod we obtain . ∎
If we have two spin frames adapted to to a common geodesic, and apply a homotopy of isometries to them, for , starting from the identity , the complex distance between the spin frames remains constant; such a homotopy describes a point of the universal cover . Hence complex distance between spin frames is invariant under the action of . 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 . In other words, if and are spin-decorated horospheres, then the complex lambda length from to is equal to the complex lambda length from to .
We can now prove Theorem 2: given spinors , and corresponding spin-decorated horospheres , the complex lambda length form to satisfies
Proof of Theorem 2.
Recalling that the spinor corresponds to a horosphere with centre , we observe that are linearly dependent (over precisely when have common centre. In other words, precisely when . We can thus assume are linearly independent.
First we prove the result when and . From Proposition 3.9 then is a spin lift of the decorated horosphere centred at with height and decoration specified by ; and is a spin lift of the decorated horosphere centred at with Euclidean diameter and decoration north-pole specified by . They are thus tangent at the point , at which point and project to coincident frames, hence either coincide or differ by .
To see that they coincide, we consider the following matrix , regarded as the lift to the universal cover of the path for , starting at the identity:
Clearly , so by -equivariance . Geometrically, in the upper half space model, is a rotation of angle about the oriented geodesic from to . Over , the point and the vector specifying the decorations remain fixed, and the frame at rotates by about to arrive at . Applying Definition 4.3, we then obtain the associated outward spin frame by a rotation of about the decoration vector, i.e. about the same axis . Thus indeed , their complex distance is , and .
Next we prove the result when and for some complex . In this case is the spin lift of a decorated horosphere centred at , with Euclidean diameter and decoration north-pole specified by . The common perpendicular runs from to , intersecting at and at . Thus the signed translation distance from to is and the rotation angle is given by mod ; lifting to spin frames we show it is indeed mod . Consider again an lifting a path from the identity,
where we take and . We have , so sends (i.e. from the previous case) to here. Geometrically is a translation of length and rotation of angle about , so as a spin isometry translates by and rotates by modulo . Since the complex distance from to at was zero in the previous case , the complex distance now becomes mod . Thus .
Finally, we prove the result for general linearly independent . There exists such that and , where . Applying this then the complex lambda length from to is equal to the complex lambda length from to , which is 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 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 , and a spin decoration on each ideal vertex . We must show that the complex lambda lengths from to satisfy
(5.1) |
Proof of Theorem 3.
Let be the spinor corresponding to . Let be the complex matrices whose ’th column is , and let be the submatrix whose columns are and in order. Then , so the claimed equation becomes
which is a well known Plücker relation. ∎
Note that if we multiply any one of the spinors, say corresponding to , by a complex scalar , each term of the Ptolemy equation (5.1) involving index is also scaled by . For instance if we multiply by then are all multiplied by . 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
(5.2) |
relating shape parameters for a hyperbolic ideal tetrahedron , as we see next. By the shape parameter of along an edge , we mean the complex number such that, if we place the two endpoints of at and , and place the remaining two ideal vertices at and a point with positive imaginary part, then the final vertex lies at . By this definition, opposite edges of have the same shape parameter, and the three pairs of shape parameters can be denoted such that and . In particular, (5.2) holds, and continues to hold if we cyclically permute .
Proposition 5.3.
Numbering the ideal vertices of by as in Figure 5, let the shape parameter of edge by . Choose a spin-decorated horosphere at ideal vertex and let be the complex lambda length from to . Then
(5.4) |
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 respectively at respectively, so . With this arrangement then , and . If we multiply a spinor corresponding to by a complex scalar , 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 , , , . By Theorem 2 then we calculate all complex lambda lengths as
This immediately gives the first equation. Permuting labels on , similarly permuting indices on each , and using the antisymmetry of , then gives the subsequent two equations. ∎
When the ideal tetrahedron is degenerate and lies in a plane, by an isometry we may place on the hyperbolic plane with , i.e. the upper half plane model inside the upper half space model. All vertices at infinity then lie in , and we may choose all spin directions to point perpendicular to in the following sense. Note that every horocycle centred at extends to a unique horosphere in , also centred at .
Definition 5.5.
Let be a horocycle in . A planar spin decoration on is a spin decoration on such that project to frames specified by .
“Specified” here means north-pole specified, if 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 will be a planar spin decoration on a horocycle if and only if the centre and the decoration direction (if ) or (if ) is a positive real multiple of . This happens precisely when are real. ∎
Thus, we can reduce to two dimensions by considering real spinors, i.e. those in . Then all complex distances between horospheres are real, so the are positive, giving Penner’s real lambda lengths between horocycles from [16, 17].
A horocycle in has two planar spin decorations, corresponding to the two spin lifts of frames specified by the 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 via the upper half plane model, but to have the orientation induced by the normal vector in the direction in the upper half space model; this is the opposite to the usual orientation. Then the boundary circle obtains an orientation in the negative real direction.
Definition 6.1.
An ideal -gon is a collection of distinct points in , labelled in order around the oriented boundary . A decoration on an ideal -gon is a choice of horocycle at each .
We can join the points (and back to ) successively by geodesics to form an ideal -gon in the usual sense, but we just need the . Given the orientation on , this means that the satisfy either
(6.2) |
or
(6.3) |
for some .
In 3 dimensions, we lose the ordering on vertices of an ideal -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 -gon is a collection of distinct points . A spin decoration on a skew ideal -gon is a choice of spin-decorated horosphere centred at each .
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.
-
(i)
The decorated Teichmüller space of ideal -gons is the space of all decorated ideal -gons, up to orientation-preserving isometries of .
-
(ii)
The decorated Teichmüller space of skew ideal -gons is the space of all spin-decorated skew ideal -gons, up to spin isometries of .
In other words, the orientation-preserving isometry group acts on the space of decorated ideal -gons, and is the quotient. Similarly, the spin isometry group acts on the space of spin-decorated skew ideal -gons, and 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 is totally positive if they are all real, and for all we have .
Note that the totally positive condition implies that the are all linearly independent, so the corresponding horospheres are all centred at distinct points ; and by antisymmetry, when we have .
Lemma 6.7.
Let . If are totally positive then the centres of the corresponding horospheres form an ideal -gon. The planar spin-decorated horospheres of yield a map
which is surjective and 2-1, with the preimage of each ideal -gon of the form .
In other words, the totally positive condition forces the to be in order around , and we obtain a decorated ideal -gon. Conversely, any decorated ideal -gon is realised by precisely two -tuples of totally positive real spinors, which are negatives of each other.
Proof.
Letting be totally positive we have
(6.8) |
Supposing then, we have , so and have the same sign when , and and have opposite signs when .
If are real and satisfy , or , or , then we obtain a contradiction. We show this in the case ; the other cases are similar. From , and have opposite signs. From , and have opposite signs. Thus and have the same sign. But and have opposite signs since , a contradiction.
This argument applies not just to but to any such that . These are precisely the cases in which are not in order around . Thus every triple of real numbers among the is in order around .
If all are real, then considering the triples , , , shows that all are in order around , satisfying (6.2).
Suppose some , so . We then have . For then is positive, so has the same sign as . Similarly, for , has opposite sign to . Thus if then have opposite signs, so from (6.8) then . Applying the reasoning of the previous paragraph, it follows that (6.3) is satisfied.
Thus the are in order around and form an ideal -gon, and by Lemma 5.6 each yields a planar spin-decorated horosphere at , hence a horocycle decoration in .
Conversely, suppose the with horocycles form a decorated ideal -gon in . Each has two planar spin decorations, given by two real spinors of the form . Choosing a sign for , the requirement that each forces a choice for each other . This yields two possible -tuples of real spinors describing the decorated ideal -gon; we will show they are both totally positive.
We can then give a description of in terms of totally positive spinors. The action of on real spinors extends to an action on -tuples as . We then obtain the following.
Proposition 6.9.
Let . The -orbits of totally positive -tuples of real spinors are naturally bijective with :
(6.10) |
Proof.
We first show the map of Lemma 6.7, sending totally positive -tuples to decorated ideal -gons, descends to a map as in (6.10). If two totally positive -tuples are related by the action of , then by equivariance of the action of , the resulting spin-decorated horospheres are related by a spin isometry of , and dropping spin structures, the underlying decorated ideal -gons are related by an isometry in .
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 -gon being negatives of each other. These two preimages are are related by the action of the negative identity in , 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 over a field is the space of all -planes in . It can be realised as the quotient of , the space of all matrices over of rank , by the left action of . A matrix represents the -plane spanned by its rows. The minors of a matrix yield projective coordinates on called Plücker coordinates. We only consider and or .
Definition 6.11.
Let denote the space of all real matrices with all Plücker coordinates positive. The positive Grassmannian is the quotient of by the left action of . The positive affine Grassmannian is the affine cone on .
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 , as in [7, example 12.6], can be identified with the nonzero decomposable elements of . The plane spanned by two rows in a matrix is represented by , and the , and the action of is by
Taking the quotient by thus identifies matrices whose corresponding decomposable elements of represent the same -plane in . Taking the quotient by identifies matrices whose corresponding elements of are equal, and thus the affine cone on is the quotient of by the left action of . Restricting to matrices in , taking the quotient by again identifies matrices whose corresponding decomposable elements of which represent the same -plane, and taking the quotient by identifies matrices whose corresponding elements of are equal. Thus is the quotient of by the left action of .
Proof of Theorem 4(i).
We now have, by Proposition 6.9 and the above discussion
Placing a -tuple of real spinors as the columns of a matrix, the totally positive condition is that for . Each such is then none other than the determinant of the minor formed by columns and , i.e. the Plücker coordinate , so we precisely obtain the matrices in . The actions of on totally positive -tuples and are identical, so we obtain an identification . By Theorem 2 each (complex) lambda length on is equal to , which we have seen is equal to the Plücker coordinate on . ∎
In a similar fashion over , we can consider the subvariety of the Grassmannian where all Plücker coordinates are nonzero.
Definition 6.12.
Let denote the space of all complex matrices with all Plücker coordinates nonzero. The nonzero Grassmannian is the quotient of by the left action of . The nonzero affine Grassmannian is the affine cone on .
Again the affine cone on can be identified with nonzero decomposable elements in , and taking the quotient by identifies precisely those matrices whose corresponding elements of are equal. Thus is the quotient of by the left action of .
Proof of Theorem 4(ii).
In a spin-decorated skew ideal -gon, at each ideal vertex we have a spin-decorated horosphere corresponding to a spinor . The fact that all are distinct (Definition 6.4) implies that for all we have . By Definition 6.5, is the space of all spin-decorated skew ideal -gons, up to spin isometries, so
Again, putting the spinors as the columns of a matrix and noting that the actions are identical gives an identification , and each complex lambda length is equal to the corresponding Plücker coordinate . ∎
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] , -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 -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 -polynomial, Math. Z. 283 (2016), no. 1-2, 515–537.