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

Gonality and genus of canonical components of character varieties

Kathleen L. Petersen
Alan W. Reid

1. Introduction

Throughout the paper, MM will always denote a complete, orientable finite volume hyperbolic 3-manifold with cusps. By abuse of notation we will denote by M\partial M to be the boundary of the compact manifold obtained from MM by truncating the cusps.

Given such a manifold, the SL2()\text{SL}_{2}(\mathbb{C}) character variety of MM, X(M)X(M), is a complex algebraic set associated to representations of π1(M)SL2()\pi_{1}(M)\rightarrow\text{SL}_{2}(\mathbb{C}) (see §4 for more details). Work of Thurston showed that any irreducible component of such a variety containing the character of a discrete faithful representation has complex dimension equal to the number of cusps of MM. Such components are called canonical components and are denoted X0(M)X_{0}(M). Character varieties have been fundamental tools in studying the topology of MM (we refer the reader to [23] for more), and canonical components carry a wealth of topological information about MM, including containing subvarieties associated to Dehn fillings of MM.

When MM has exactly one cusp, any canonical component is a complex curve. The aim of this paper is to study how some of the natural invariants of these complex curves correspond to the underlying manifold MM. In particular, we concentrate on how the gonality of these curves behaves in families of Dehn fillings on 2-cusped hyperbolic manifolds. More precisely, we study families of 1-cusped 3-manifolds which are obtained by Dehn filling of a single cusp of a fixed 2-cusped hyperbolic 3-manifold, MM. We write M(,r)M(-,r) to denote the manifold obtained by r=p/qr=p/q filling of the second cusp of MM.

To state our results we introduce the following notation. If XX is a complex curve, we write γ(X)\gamma(X) to denote the gonality of XX, g(X)g(X) to be the (geometric) genus of XX and d(X)d(X) to be the degree (of the specified embedding) of XX. The gonality of a curve is the lowest degree of a map from that curve to \mathbb{C}. Unlike genus, gonality is not a topological invariant of curves, but rather is intimately connected to the geometry of the curve. There are connections between gonality and genus, most notably the Brill-Noether theorem which gives an upper bound for gonality in terms of genus (see § 7) but in some sense these are orthogonal invariants. For example, all hyperelliptic curves all have gonality two, but can have arbitrarily high genus. Moreover, for g>2g>2, there are curves of genus gg of different gonality. We refer the reader to § 3 for precise definitions.

Our first theorem is the following.

Theorem 1.1.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, then there is a positive constant cc depending only on MM such that

γ(X0(M(,r)))c.\gamma(X_{0}(M(-,r)))\leq c.

The figure-8 knot complement and its so-called sister manifold are both integral surgery of one component of the Whitehead link complement. Their canonical components are well-known to be an elliptic curve, and a rational curve, respectively. These have gonality 2 and 1. Therefore, Theorem 1.1 is best possible.

Our techniques can also be used to obtain information about the genera and degree of related varieties as we now discuss. The inclusion map from π1(M)\pi_{1}(\partial M) to π1(M)\pi_{1}(M) induces a map from X(M)X(M) to the character variety of M\partial M. We let A(M)A(M) denote the image of this map and let A0(M)A_{0}(M) denote the image of a canonical component. When MM has two cusps, the variety A(M)A(M) naturally sits in 4(m1,l1,m2,l2)\mathbb{C}^{4}(m_{1},l_{1},m_{2},l_{2}) where the mim_{i} and lil_{i} are a choice of framing (we often refer to these as meridional and longitudinal parameters for the ithi^{th} cusp). For r=p/q0r=p/q\neq 0 in lowest terms, we define the naive height of rr to be h(r)=max{|p|,|q|}h(r)=\max\{|p|,|q|\}. To avoid cumbersome notation, we define h(0)=h()=1h(0)=h(\infty)=1.

Theorem 1.2.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, then there is a positive constant cc depending only on MM and the framing of the second cusp such that

g(A0(M(,r)))ch(r)2.g(A_{0}(M(-,r)))\leq c\cdot h(r)^{2}.

Using work of Dunfield [7] (see also §4) we deduce

Corollary 1.3.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, and |H1(M(,r);/2)|=2|H^{1}(M(-,r);\mathbb{Z}/2\mathbb{Z})|=2 then there is a positive constant cc depending only on MM and the framing of the second cusp such that

g(X0(M(,r)))ch(r)2.g(X_{0}(M(-,r)))\leq c\cdot h(r)^{2}.

Unlike the genus or gonality of a curve, the degree of a variety is inherently tied to its embedding in ambient space. We obtain upper bounds for the degree of a ‘natural model’ of A0(M(,r))A_{0}(M(-,r)).

The proof of Theorem 1.1 breaks naturally into cases, depending on whether or not one (or both) cusps of MM are geometrically isolated from the other cusp (see [17] and § 5). Geometric isolation also plays a role in our determination of degree bounds.

Theorem 1.4.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, then there is a positive constant c1c_{1} depending only on MM and the framing of the second cusp such that

d(A0(M(,r)))c1h(r).d(A_{0}(M(-,r)))\leq c_{1}\cdot h(r).

If one cusp is geometrically isolated from the other cusp, then there is a positive constant c2c_{2} depending only on MM and the framing of the second cusp, such that

d(A0(M(,r)))c2.d(A_{0}(M(-,r)))\leq c_{2}.

Work of Hoste and Shanahan [11] demonstrates that the degree bounds in Theorem 1.4 are sharp.

In contrast with the main results of this paper, we also show that by filling two cusps of a three cusped manifold, we can construct examples of knots (so-called double twist knots) for which the gonality of the canonical component can be made arbitrarily large (see §11.2 for details). Existence of families of curves with large gonality has been the subject of much recent research, including connections between the existence of towers of curves whose gonality goes to infinity and expander graphs [8].

We also remark on the gonalities of the character varieties of the once-punctured torus bundles of tunnel number one [1]. These manifolds are all integral surgeries on one cusp of the Whitehead link and the corresponding gonalities are all equal to one or two. Both the double twist knot examples and the once-punctured torus bundles of tunnel number one examples illustrate linear growth in the genus of the character varieties. We do not know if a quadratic bound such as in Theorem 1.2 is optimal.

2. Acknowledgements

The first author would like to thank the University of Texas at Austin and the Max Planck Institut für Mathematik for their hospitality while working on this manuscript. She would also like to thank Martin Hils for helpful discussions. This work was partially supported by a grant from the Simons Foundation (#209226 to Kathleen Petersen), and by the National Science Foundation (to Alan Reid).

3. Algebraic Geometry and Preliminary Lemmas

3.1. Preliminaries

We begin with some algebraic geometry. We follow Shafarevich [22] and Hartshorne [9]. The ‘natural models’ of the character varieties of interest are complex affine algebraic sets, which typically have singularities at infinity. We will rely on an analysis of rational maps between possibly singular affine sets. Such a map f:XYf:X\dashrightarrow Y behaves (on most of XX and f(Y)f(Y)) as a combination of a very nice map and a map with controlled bad parts. We now present some algebraic geometry to make this precise.

Let [T]\mathbb{C}[T] be the polynomial ring with coefficients in \mathbb{C} in the variables T1,,TnT_{1},\dots,T_{n}. A (Zariski) closed subset of n\mathbb{C}^{n} is a subset XnX\subset\mathbb{C}^{n} consisting of the common vanishing set of a finite number of polynomials in [T]\mathbb{C}[T]. We let 𝔘X\mathfrak{U}_{X} denote the ideal of XX, the ideal in [T]\mathbb{C}[T] consisting of all polynomials which vanish on XX. If II is an ideal in [T]\mathbb{C}[T] we let V(I)V(I) denote the vanishing set of II in n\mathbb{C}^{n}, and if f[T]f\in\mathbb{C}[T] let V(f)V(f) be the vanishing set of ff in n\mathbb{C}^{n}. The coordinate ring of XX is [X]=[T]/𝔘X\mathbb{C}[X]=\mathbb{C}[T]/{\mathfrak{U}}_{X}. This ring is the ring of all polynomials on XX, up to the equivalence of being equal on XX; adding any polynomial in 𝔘X{\mathfrak{U}}_{X} to a polynomial ff does not change the value of ff on XX. If a closed set XX is irreducible then the field of fractions of the coordinate ring [X]\mathbb{C}[X] is the function field or field of rational functions of XX; it is denoted (X)\mathbb{C}(X).

A function ff defined on XX with values in \mathbb{C} is regular if there exists a polynomial F(T)F(T) with coefficients in \mathbb{C} such that f(x)=F(x)f(x)=F(x) for all xXx\in X. A map f:XYf:X\rightarrow Y is regular if there exist nn regular functions f1,,fnf_{1},\dots,f_{n} on XX such that f(x)=(f1(x),,fn(x))f(x)=(f_{1}(x),\dots,f_{n}(x)) for all xXx\in X. Any regular map f:XYf:X\rightarrow Y defines a pullback homomorphism of \mathbb{C}-algebras f:[Y][X]f^{*}:\mathbb{C}[Y]\rightarrow\mathbb{C}[X] as follows. If u:XZu:X\rightarrow Z we define v:YZv:Y\rightarrow Z by v(x)=u(f(x))v(x)=u(f(x)). We define ff^{*} by f(u)=vf^{*}(u)=v. A rational map f:XYmf:X\dashrightarrow Y\subset\mathbb{C}^{m} is an mm-tuple of rational functions f1,,fm(X)f_{1},\dots,f_{m}\in\mathbb{C}(X) such that, for all points xXx\in X at which all the fif_{i} are defined, f(x)=(f1(x),,fm(x))Y.f(x)=(f_{1}(x),\dots,f_{m}(x))\in Y. We define the pullback ff^{*} as above. A rational map f:XYf:X\dashrightarrow Y is called dominant if f(X)f(X) is (Zariski) dense in YY. A rational map f:XYf:X\dashrightarrow Y is called birational if ff has a rational inverse, and in this case we say that XX and YY are birational. A regular map f:XYf:X\rightarrow Y is finite if [X]\mathbb{C}[X] is integral over [Y]\mathbb{C}[Y]. These definitions naturally extend to the projective setting. A morphism φ:XY\varphi:X\rightarrow Y is a continuous map such that for every open VYV\subset Y and every regular function f:Yf:Y\rightarrow\mathbb{C} the function fφ:φ1(Y)f\circ\varphi:\varphi^{-1}(Y)\rightarrow\mathbb{C} is regular.

Character varieties are naturally affine sets. It is useful to consider an embedding of such an affine set into projective space. Given an affine variety XX, a projective completion of XX is the Zariski closure of an embedding of XX in some projective space. A smooth projective model of XX is a smooth projective variety birational to a projective completion of XX. Such a model is unique up to isomorphism if XX is a curve. The degree of any projective curve XX, denoted d(X)d(X) is the maximum number of points of intersection of XX with a hyperplane not containing any component of XX. This depends on the specific embedding of XX.

The gonality of a curve is, in the most basic sense, the degree of a map from that curve to 1\mathbb{C}^{1}. Intuitively the degree of a map is the number of preimages of ‘most points’ in the image. Let XX and YY be irreducible varieties of the same dimension and f:XYf:X\dashrightarrow Y a dominant rational map. The degree of the field extension f((Y))(X)f^{*}(\mathbb{C}(Y))\subset\mathbb{C}(X) is finite and is called the degree of ff. That is,

degf=[(X):f((Y))].\text{deg}f=[\mathbb{C}(X):f^{*}(\mathbb{C}(Y))].

This is well defined for rational and regular maps.

We are now in a position to see that dominant rational maps between possibly singular curves can be decomposed in a pleasing manner.

Proposition 3.1.

Let XX be a possibly singular irreducible affine variety of dimension nn and YY an algebraic set of dimension nn. Assume that g:XYg:X\dashrightarrow Y is a dense rational map.

  1. (1)

    deg(g)=d\deg(g)=d is finite.

  2. (2)

    For all but finitely many co-dimension one subvarieties WW of XX,

    deg(gW)d.\deg(g\mid_{W})\leq d.
  3. (3)

    If n2n\leq 2 then g1(y)g^{-1}(y) consists of at most dd points except for perhaps finitely many yYy\in Y. In XX there are finitely many irreducible sets of co-dimension one such that if xx is not contained in the union of these sets then g1(g(x))g^{-1}(g(x)) contains only finitely many points.

Proof.

By Hironaka [10] one can resolve the singularities of a variety by a proper birational map (which has degree one). Therefore, the variety XX is birational to a smooth projective variety XX^{\prime}, and similarly YY is birational to a smooth projective YY^{\prime}. The rational map g:XYg:X\dashrightarrow Y induces a rational map g:XYg^{\prime}:X^{\prime}\dashrightarrow Y^{\prime} of the same degree by pre and post composition with the birational inverses,

X𝛼XgY𝛽Y.X\overset{\alpha}{\dashrightarrow}X^{\prime}\overset{g^{\prime}}{\rightarrow}Y^{\prime}\overset{\beta}{\dashrightarrow}Y.

By Hironaka’s resolution of indeterminacies [10], gg^{\prime} can be resolved into a regular map by a sequence of blowups. Therefore we can take gg^{\prime} to be a proper morphism. By Stein Factorization (see [9] Corollary 11.5), a proper morphism factors as a proper morphism with connected fibers followed by a finite map. It follows that the degree is finite. By the Weak Factorization Theorem [26] these birational maps (since they are over \mathbb{C}) can be decomposed as a series of blow-ups and blow-downs over smooth complete varieties.

A blow up or blow down is defined for all points outside of a codimension one set. Therefore, outside of a finite union of codimension one sub varieties, the map gg is the composition of isomorphisms, a proper morphism with connected fibers, and a finite map. If WXW\subset X is a codimension one subvariety not in this set, then the degree of gWg\mid_{W} is at most the degree of gg.

If XX is a curve, then part 3 follows from the observation that in this case, Stein Factorization implies that gg^{\prime} is a finite map. If XX is a surface, then it suffices to show content of the lemma for XX and YY smooth projective surfaces, and g:XYg:X\rightarrow Y a dense morphism. By Stein factorization such a gg can be decomposed into a map with connected fibers followed by a finite map. For the finite map, the number of pre-images of a point is bounded above by the degree. Since the maps are between surfaces, the map with connected fibers must be birational, so part 3 follows.

Lemma 3.2.

Let f:XYf:X\dashrightarrow Y be a dominant rational map of affine algebraic sets. If XX is irreducible, then YY is irreducible. (That is, the closure of f(X)f(X) is irreducible.)

Proof.

By Proposition 3.1 it suffices to assume that ff is regular. The map ff induces a map f:[Y][X]f^{*}:\mathbb{C}[Y]\rightarrow\mathbb{C}[X] of coordinate rings. If X[T]X\subset\mathbb{C}[T] and Y[S]Y\subset\mathbb{C}[S] then [X]=[T]/𝔘X\mathbb{C}[X]=\mathbb{C}[T]/{\mathfrak{U}}_{X} and [Y]=[S]/𝔘Y\mathbb{C}[Y]=\mathbb{C}[S]/{\mathfrak{U}}_{Y}. Since f(X)f(X) is dense in YY, it follows that ff^{*} is an isomorphic inclusion (see [22] page 31). An algebraic set is irreducible if and only if the vanishing ideal is prime. Therefore 𝔘X{\mathfrak{U}}_{X} is prime and [X]\mathbb{C}[X] has no zero divisors. We conclude that [Y]\mathbb{C}[Y] has no zero divisors, and therefore 𝔘Y{\mathfrak{U}}_{Y} is prime and YY is irreducible.

3.2. Gonality

The algebraic definition of gonality is as follows.

Definition 3.3.

The gonality of a curve XX in n\mathbb{P}^{n} is

γ(X)=min{[(X):(x)]:x(X)}.\gamma(X)=\min\{[\mathbb{C}(X):\mathbb{C}(x)]:x\in\mathbb{C}(X)-\mathbb{C}\}.

Here, x(X)x\in\mathbb{C}(X) is generated by a single element. Since (X)\mathbb{C}(X) is the function field of XX and consists of the rational functions on XX, we can reformulate the definition; the gonality of XX is the minimal degree of a dominant rational map from XX to 1\mathbb{P}^{1}. Therefore, gonality is a birational invariant; two varieties which are birational (but not necessarily isomorphic) have equal gonality. If XX is an affine curve, XX is birational to a smooth projective model and as such the gonality of XX and XX^{\prime} are equal, and this is the minimal degree of a dominant rational map from XX to \mathbb{C}.

By Noether’s Normalization theorem, if XX is complex projective variety of dimension nn, then there is a linear subspace disjoint from XX such that the projection onto that subspace is finite to one and surjective. This is also true for affine varieties; if ZZ is an affine variety of dimension nn there is a finite degree map from ZZ onto a dense subset of n\mathbb{C}^{n}. Therefore, gonality is always finite.

Remark 3.4.

Gonality of an affine variety is sometimes defined as the minimal degree of a regular map to \mathbb{C} (or of a finite map to \mathbb{C}). This (regular) gonality is easily seen to be greater than or equal to the (rational) gonality. In fact, there are examples where these two notions differ. The affine curve determined by the equation xy=1xy=1 is not isomorphic to \mathbb{C} and so the (regular) gonality is not one. One can see that there is a regular degree two map onto \mathbb{C}, by projecting onto the line y=xy=-x. Therefore, the (regular) gonality is two. Projection onto the xx coordinate is a dominant map to \mathbb{C} as the image is {0}\mathbb{C}-\{0\} and has degree one; the (rational) gonality is one.

In the projective setting the gonality defined by using rational, regular, or finite maps are all equal, and are therefore equal to the (rational) gonality of a corresponding affine variety. If there is a dominant rational map g:XYg:X\dashrightarrow Y between affine (possibly singular) curves then this induces a finite dominant morphism ff between smooth projective models XX^{\prime} and YY^{\prime} for XX and YY, where the degree of gg and the degree of ff are the same.

Gonality can also be defined using the language of Riemann surfaces instead of complex curves. Here the definitions involving rational or regular maps correspond to meromorphic or holomorphic functions from the Riemann surface to 1\mathbb{P}^{1}. Similarly, one can define the gonality in terms of the minimal degree of a branched or unbranched covering of 1\mathbb{P}^{1} by the Riemann surface.

3.3. Key Lemmas

The following lemma will be essential for our gonality bounds.

Lemma 3.5.

Let g:XYg:X\dashrightarrow Y be a degree dd dominant rational map of irreducible (possibly singular) affine or projective curves. Then

γ(Y)γ(X)dγ(Y).\gamma(Y)\leq\gamma(X)\leq d\cdot\gamma(Y).
Proof.

The second inequality is clear. To show the first, we follow [20] Proposition A.1, including the proof for completeness. Let f(X)f\in\mathbb{C}(X) be a degree dd map realizing the gonality of XX. Let p(T)(Y)[T]p(T)\in\mathbb{C}(Y)[T] be the characteristic polynomial of ff when ff is viewed as an element in the extension (X)\mathbb{C}(X) of (Y)\mathbb{C}(Y). Let MM be a finite extension of (Y)\mathbb{C}(Y) such that p(T)p(T) factors into the product of rr monomials (Tfi)(T-f_{i}), with fiMf_{i}\in M.

Viewed as a function in MM, the degree of ff is d[M:(X)]d[M:\mathbb{C}(X)]. This is also the degree of the fif_{i} since they are all in the same Aut(M/(Y)){\rm Aut}(M/\mathbb{C}(Y)) orbit. The polar divisors of a coefficient of pp viewed in MM is at most the sum of the polar divisors of the fif_{i}. (These polar divisors are just the effective divisors (fi)(f_{i})_{\infty}.)

Therefore, each coefficient has degree at most dr[M:(X)]=d[M:(Y)]dr[M:\mathbb{C}(X)]=d[M:\mathbb{C}(Y)] as a function in MM, and therefore degree at most dd as a function in (Y)\mathbb{C}(Y). Since ff is non-constant, at least one of these coefficients is non-constant. Therefore the gonality of YY is at most dd.

We now introduce some notation. Commonly α(x)Θ(β(x))\alpha(x)\in\Theta(\beta(x)) is written to mean that there are positive constants c1c_{1} and c2c_{2} independent of xx such that

c1β(x)α(x)c2β(x).c_{1}\beta(x)\leq\alpha(x)\leq c_{2}\beta(x).

Notice that if α(x)Θ(β(x))\alpha(x)\in\Theta(\beta(x)) then also

c21α(x)β(x)c11α(x)c_{2}^{-1}\alpha(x)\leq\beta(x)\leq c_{1}^{-1}\alpha(x)

so β(x)Θ(α(x))\beta(x)\in\Theta(\alpha(x)). This defines an equivalence relation.

Definition 3.6.

For positive quantities α=α(x)\alpha=\alpha(x) and β=β(x)\beta=\beta(x) we write αβ\alpha\sim\beta if αΘ(β)\alpha\in\Theta(\beta).

We will use \sim in conjunction with quantities such as γ(X0(M(,r)))\gamma(X_{0}(M(-,r))) to mean that the constants do not depend on rr but only on MM and the framing of M\partial M. Note that α(x)1\alpha(x)\sim 1 means that there is a positive constant cc independent of xx such that α(x)c\alpha(x)\leq c.

We now compute the gonality of the curves given by xpyqx^{p}-y^{q} and xpyq1x^{p}y^{q}-1 in 2(x,y)\mathbb{C}^{2}(x,y). First we establish some notation for these sets.

Definition 3.7.

Let pp and qq be relatively prime integers with q>0q>0. With r=p/qr=p/q, let

𝒞(r)=V(x|p|yq1)V(x|p|yq)\mathcal{C}(r)=V(x^{|p|}y^{q}-1)\cup V(x^{|p|}-y^{q})

in 2(x,y)\mathbb{C}^{2}(x,y) and let

(r)=V(x|p|yq1)V(x|p|yq)\mathcal{H}(r)=V(x^{|p|}y^{q}-1)\cup V(x^{|p|}-y^{q})

in 4\mathbb{C}^{4} where xx and yy are among the coordinates. (Therefore, (r)\mathcal{H}(r) and 𝒞(r)\mathcal{C}(r) are each the union of two hypersurfaces unless p=0p=0.)

Lemma 3.8.

Assume that pp and qq are relatively prime integers. Then the gonality of each irreducible component of 𝒞(r)2(x,y)\mathcal{C}(r)\subset\mathbb{C}^{2}(x,y) is one.

Proof.

If either pp or qq is zero, the polynomial is x=1x=1 or y=1y=1 and is therefore naturally a 1\mathbb{C}^{1} and has gonality one. Now assume that p,q>0p,q>0. We will compute the gonality of V(xpyq1)V(x^{p}y^{q}-1) as the vanishing set of xpyqx^{p}-y^{q} since this vanishing set is birationally equivalent to the vanishing set of xpyq1x^{p}y^{q}-1 under the map (x,y)(x,y1)(x,y)\mapsto(x,y^{-1}).

Since pp and qq are relatively prime, there are integers aa and bb such that ap+bq=1ap+bq=1. Let φ:2(x,y)2(x,y)\varphi:\mathbb{C}^{2}(x,y)\dashrightarrow\mathbb{C}^{2}(x^{\prime},y^{\prime}) be the rational map given by

(x,y)(xbya,xpyq)=(x,y).(x,y)\mapsto(x^{b}y^{-a},x^{p}y^{q})=(x^{\prime},y^{\prime}).

This has an inverse φ1\varphi^{-1} given by

(x,y)((x)q(y)a,(x)p(y)b).(x^{\prime},y^{\prime})\mapsto((x^{\prime})^{q}(y^{\prime})^{a},(x^{\prime})^{-p}(y^{\prime})^{b}).

The map φ\varphi is birational, so γ(V(xpyq1))\gamma(V(x^{p}y^{q}-1)) is equal to the gonality of the image of V(xpyq1)V(x^{p}y^{q}-1) under φ\varphi. This image is dense in y=1y^{\prime}=1. As y=1y^{\prime}=1 naturally defines a 1\mathbb{C}^{1} the gonality of V(y1)V(y^{\prime}-1) and V(xpyq1)V(x^{p}y^{q}-1) are both one.

3.4. Gonality and Projection

We will now prove a few lemmas which will be essential in bounding the gonalities we are interested in. In §6\S~\ref{section:gonalityindehnsurgeryspace} we will relate the situation below to that of the character varieties of interest. Recall the definition of 𝒞(r)\mathcal{C}(r) and (r)\mathcal{H}(r) from Definition 3.7. By Lemma 3.8 the gonality of (each irreducible component of) 𝒞(r)2\mathcal{C}(r)\subset\mathbb{C}^{2} is equal to one.

Remark 3.9.

If a curve CC is reducible, by γ(C)\gamma(C) we mean the maximal gonality of any irreducible curve component of CC.

Definition 3.10.

Define the maps ϖ1\varpi_{1} and ϖ2\varpi_{2} to be the projection maps

ϖi:4(x1,y1,x2,y2)2(xi,yi)\varpi_{i}:\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2})\rightarrow\mathbb{C}^{2}(x_{i},y_{i})

for i=1,2i=1,2.

By Definition 3.7, with (r)\mathcal{H}(r)\subset 4(x1,y1,x2,y2)\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2}) and 𝒞(r)2(x2,y2)\mathcal{C}(r)\subset\mathbb{C}^{2}(x_{2},y_{2}) then 𝒞(r)=ϖ2((r)).\mathcal{C}(r)=\varpi_{2}(\mathcal{H}(r)).

Lemma 3.11.

Let VV be an irreducible surface in 4(x1,y1,x2,y2)\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2}) with dimϖi(V)>0\dim_{\mathbb{C}}\varpi_{i}(V)>0. Then V(r)V\cap\mathcal{H}(r) consists of a finite union of curves for all but finitely many rr.

Proof.

The dimension of VV is two, and the dimension of (r)\mathcal{H}(r) is three. Therefore, each irreducible component of their intersection has dimension at least one (see [9] Section I.7). As a result, the dimension of any irreducible component of V(r)V\cap\mathcal{H}(r) is one, or two.

Assume there are infinitely many rr such that V(r)V\cap\mathcal{H}(r) contains an irreducible component of dimension two. The set (r)\mathcal{H}(r) is of the form 2×𝒞(r)\mathbb{C}^{2}\times\mathcal{C}(r), and in 2(x2,y2)\mathbb{C}^{2}(x_{2},y_{2}) any two distinct 𝒞(r)\mathcal{C}(r) intersect in only finitely many points. We conclude that V2×PV\subset\mathbb{C}^{2}\times P for some point PP. Therefore, ϖ2(V)=P\varpi_{2}(V)=P has dimension zero, which does not occur by hypothesis. It follows that V(r)V\cap\mathcal{H}(r) consists of a finite union of curves for all but finitely many rr.

Lemma 3.12.

Let VV be an irreducible surface in 4(x1,y1,x2,y2)\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2}) with dimϖi(V)>0\dim_{\mathbb{C}}\varpi_{i}(V)>0. Assume that V(r)V\cap\mathcal{H}(r) is a finite union of curves and ϖ2(V)\varpi_{2}(V) is dense in 2\mathbb{C}^{2}. Then for any irreducible curve component CC of V(r)V\cap\mathcal{H}(r)

γ(C)1\gamma(C)\sim 1

where the implied constant depends only on VV.

Proof.

By Proposition 3.1 the image ϖ2(V)\varpi_{2}(V) consists of all of 2(x2,y2)\mathbb{C}^{2}(x_{2},y_{2}) except perhaps a finite union of curves and points. The image ϖ2((r))=𝒞(r)2(x2,y2)\varpi_{2}(\mathcal{H}(r))=\mathcal{C}(r)\subset\mathbb{C}^{2}(x_{2},y_{2}) and since two distinct 𝒞(r)\mathcal{C}(r) intersect in only a finite set of points, we conclude that ϖ2(V(r))\varpi_{2}(V\cap\mathcal{H}(r)) maps onto a dense subset of 𝒞(r)\mathcal{C}(r) for all but perhaps finitely many rr. Therefore, it suffices to show the bound for these rr where ϖ2(V(r))\varpi_{2}(V\cap\mathcal{H}(r)) is dense in 𝒞(r)\mathcal{C}(r).

By Proposition 3.1, the degree d=deg(ϖ2V)d=\deg(\varpi_{2}\!\!\mid_{V}) is finite, and for all but a finite number of points {Pm}\{P_{m}\} in the image of ϖ2V\varpi_{2}\!\!\mid_{V}, the number of pre-images of any point is bounded by dd. For some such PmP_{m}, the pre-image of PmP_{m} is contained in a finite union of curves {Cm,n}\{C_{m,n}\} in VV. The set V(r)V\cap\mathcal{H}(r) is a union of curves and each of these curves intersects a Cm,nC_{m,n} in a finite collection of points unless Cm,nC_{m,n} coincides with a curve component of V(r)V\cap\mathcal{H}(r).

Let DD be an irreducible curve component of V(r)V\cap\mathcal{H}(r). If Cm,nC_{m,n} coincides with DD then as there are only finitely many Cm,nC_{m,n} the maximal gonality of any such DD is bounded by a constant depending only on VV. If Cm,nC_{m,n} does not coincide with DD then the degree of ϖ2D\varpi_{2}\!\!\mid_{D} is bounded by dd, which is finite and independent of rr. By Lemma 3.5 we conclude that γ(D)γ(𝒞(r))\gamma(D)\sim\gamma(\mathcal{C}(r)). By Lemma 3.8, γ(𝒞(r))=1\gamma(\mathcal{C}(r))=1, which concludes the proof.

Lemma 3.13.

Let VV be an irreducible surface in 4(x1,y1,x2,y2)\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2}) with dimϖi(V)>0\dim_{\mathbb{C}}\varpi_{i}(V)>0. Assume that V(r)V\cap\mathcal{H}(r) is a finite union of curves and ϖ2(V)\varpi_{2}(V) is a curve. Then for any irreducible curve component CC of V(r)V\cap\mathcal{H}(r)

γ(C)1\gamma(C)\sim 1

where the implied constant depends only on VV.

Proof.

The projection ϖ2(V(r))\varpi_{2}(V\cap\mathcal{H}(r)) is a finite collection of points unless a curve component of 𝒞(r)\mathcal{C}(r) is contained in the closure of ϖ2(V)2(x2,y2)\varpi_{2}(V)\subset\mathbb{C}^{2}(x_{2},y_{2}). This can only occur for finitely many rr as ϖ2(V)\varpi_{2}(V) is a curve. For all rr in this finite set, γ(V(r))\gamma(V\cap\mathcal{H}(r)) is bounded.

It now suffices to consider those rr such that ϖ2(V(r))\varpi_{2}(V\cap\mathcal{H}(r)) is a finite collection of points. For these rr, any curve in V(r)V\cap\mathcal{H}(r) must be of the form D×PD\times P where D2(x1,y1)D\subset\mathbb{C}^{2}(x_{1},y_{1}) is a curve and P2(x2,y2)P\subset\mathbb{C}^{2}(x_{2},y_{2}) is a point. (If (1,1)(1,1) is in ϖ2(V)\varpi_{2}(V) then it is possible that PP are the same for all rr since (1,1)(1,1) is contained in each DD.)

It suffices to universally bound the gonality of one such Dr×PrD_{r}\times P_{r} as there are only finitely many irreducible components of each V(r).V\cap\mathcal{H}(r). Call such a point Pr=(ar,br)P_{r}=(a_{r},b_{r}). We use the notation established in § 3.1. Let {φi}\{\varphi_{i}\} be a generating set for 𝔘V\mathfrak{U}_{V} where φi=φi(x1,y1,x2,y2)\varphi_{i}=\varphi_{i}(x_{1},y_{1},x_{2},y_{2}). A generating set for 𝔘Dr×Pr\mathfrak{U}_{D_{r}\times P_{r}} is contained in {φr,i}\{\varphi_{r,i}^{*}\} where φr,i=φi(x1,y1,ar,br)\varphi_{r,i}^{*}=\varphi_{i}(x_{1},y_{1},a_{r},b_{r}). The degrees of φix1\varphi_{i}\!\!\mid_{x_{1}} and φiy1\varphi_{i}\!\!\mid_{y_{1}} provide an upper bound for the degrees of φr,i\varphi_{r,i}^{*}. These upper bounds are independent of rr. It follows that the degree drd_{r} of Dr×PrD_{r}\times P_{r} is bounded by some dd which is independent of rr.

The genus of Dr×PrD_{r}\times P_{r} equals the genus of DrD_{r} as they are isomorphic. Moreover, the degree of DrD_{r} equals the degree of Dr×PrD_{r}\times P_{r} as seen by the form of the φr,i\varphi_{r,i}^{*}. Therefore, we identify Dr×PrD_{r}\times P_{r} with the plane curve Dr2(x1,y1)D_{r}\subset\mathbb{C}^{2}(x_{1},y_{1}). By the genus degree formula the genus grg_{r} of DrD_{r} satisfies

gr12(dr1)(dr2).g_{r}\leq\tfrac{1}{2}(d_{r}-1)(d_{r}-2).

(There is equality if the plane curve is non-singular.) The Brill-Noether bound states that for γr=γ(Dr),\gamma_{r}=\gamma(D_{r}),

γrgr+32.\gamma_{r}\leq\lfloor\frac{g_{r}+3}{2}\rfloor.

Therefore

γrgr+3214(dr1)(dr2)+2.\gamma_{r}\leq\lfloor\frac{g_{r}+3}{2}\rfloor\leq\frac{1}{4}(d_{r}-1)(d_{r}-2)+2.

The lemma follows as drdd_{r}\leq d and dd is independent of rr. ∎

4. Dehn Filling and Character Varieties

4.1.

Assume that MM is a finite volume hyperbolic 3-manifold whose boundary consists of exactly kk torus cusps. Fix a framing for each cusp; let TiT_{i} be the ithi^{th} connected component of M\partial M and let μi\mu_{i} and λi\lambda_{i} be an oriented choice of meridian and longitude for TiT_{i}. For ri=pi/qi{1/0}r_{i}=p_{i}/q_{i}\in\mathbb{Q}\cup\{1/0\}, i=1,,ki=1,\dots,k, written in lowest terms (with qq non-negative), let M(r1,,rk)M(r_{1},\dots,r_{k}) denote the manifold obtained by performing rir_{i} Dehn filling on the ithi^{th} cusp of MM. We write ri=r_{i}=- to indicate that the ithi^{th} cusp remains complete and is not filled. By Thurston’s Hyperbolic Dehn Surgery theorem, for all but finitely many values of rr in {1/0}\mathbb{Q}\cup\{1/0\} the manifold obtained by rr filling of one cusp of MM is hyperbolic.

We fix a finite presentation for Γπ1(M)\Gamma\cong\pi_{1}(M), with Mi=[μi]M_{i}=[\mu_{i}], Li=[λi]L_{i}=[\lambda_{i}] amongst the generators for i=1,,ki=1,\dots,k. If Γ\Gamma has presentation

Γ=γ1,,γn:ϱ1,ϱl\Gamma=\langle\gamma_{1},\dots,\gamma_{n}:\varrho_{1},\dots\varrho_{l}\rangle

then by van Kampen’s theorem, a presentation for

Γ(r1,,rk)π1(M(r1,,rk))\Gamma(r_{1},\dots,r_{k})\cong\pi_{1}(M(r_{1},\dots,r_{k}))

is given by

Γ(r1,,rk)=γ1,,γn:ϱ1,ϱl,w1,,wk\Gamma(r_{1},\dots,r_{k})=\langle\gamma_{1},\dots,\gamma_{n}:\varrho_{1},\dots\varrho_{l},w_{1},\dots,w_{k}\rangle

where wi=MipLiqw_{i}=M_{i}^{p}L_{i}^{q} unless r=p/q=r=p/q=- in which case it is empty.

We state the next lemma for convenience. The proof is clear.

Lemma 4.1.

Let MM be a finite volume two-cusped hyperbolic 3-manifold. Then

|H1(M(,r);/2)||H1(M;/2)||H^{1}(M(-,r);\mathbb{Z}/2\mathbb{Z})|\leq|H^{1}(M;\mathbb{Z}/2\mathbb{Z})|

for all r{1/0}r\in\mathbb{Q}\cup\{1/0\}.

4.2.

We will now define various varieties associated to the fundamental group of a 3-manifold MM (see [25].) For a finitely generated group Γ\Gamma, and a homomorphism ρ:ΓSL2(),\rho:\Gamma\rightarrow\text{SL}_{2}(\mathbb{C}), the set

Hom(Γ,SL2())={ρ:ΓSL2()}\text{Hom}(\Gamma,\text{SL}_{2}(\mathbb{C}))=\{\rho:\Gamma\rightarrow\text{SL}_{2}(\mathbb{C})\}

naturally carries the structure of an affine algebraic set defined over \mathbb{C}. This is called the representation variety, R(Γ)R(\Gamma). Given an element γΓ\gamma\in\Gamma we define the trace function Iγ:R(Γ)I_{\gamma}:R(\Gamma)\rightarrow\mathbb{C} as Iγ(ρ)=tr(ρ(γ))I_{\gamma}(\rho)=\text{tr}(\rho(\gamma)). Similarly, given a representation ρR(Γ)\rho\in R(\Gamma) we define the character of ρ\rho to be the function χρ:Γ\chi_{\rho}:\Gamma\rightarrow\mathbb{C} given by χρ(γ)=tr(ρ(γ))\chi_{\rho}(\gamma)=\text{tr}(\rho(\gamma)). The SL2()\text{SL}_{2}(\mathbb{C})-character variety of Γ\Gamma is the set of all characters,

X(Γ)={χρ:ρR(Γ)}.X(\Gamma)=\{\chi_{\rho}:\rho\in R(\Gamma)\}.

Such a set is called a natural model for the character variety, as opposed to a set isomorphic to one obtained in this manner. The functions tγ:R(Γ)X(Γ)t_{\gamma}:R(\Gamma)\rightarrow X(\Gamma) defined by tγ(ρ)=χρ(γ)t_{\gamma}(\rho)=\chi_{\rho}(\gamma) are regular maps. The coordinate ring of X(Γ)X(\Gamma) is TT\otimes\mathbb{C} where TT is the subring of all regular functions from R(Γ)R(\Gamma) to \mathbb{C} generated by 1 and the tγt_{\gamma} functions.

If Γ\Gamma is isomorphic to Γ\Gamma^{\prime}, then the sets R(Γ)R(\Gamma) and R(Γ)R(\Gamma^{\prime}) are isomorphic, as are X(Γ)X(\Gamma) and X(Γ)X(\Gamma^{\prime}). We write R(M)R(M) or X(M)X(M) to denote a specific R(π1(M))R(\pi_{1}(M)) or X(π1(M))X(\pi_{1}(M)), respectively, which is well-defined up to isomorphism. The inclusion homomorphism i:π1(M)π1(M)i:\pi_{1}(\partial M)\rightarrow\pi_{1}(M) induces a map i:X(M)X(π1(M))i^{*}:X(M)\rightarrow X(\pi_{1}(\partial M)) by χρχρπ1(M).\chi_{\rho}\mapsto\chi_{\rho\mid_{\pi_{1}(\partial M)}}. We write A(M)A(M) to denote the closure of the image of this map.

The representation variety R(M)R(M) naturally sits in 4n(g11,g12,g13,g14,,gn4)\mathbb{C}^{4n}(g_{11},g_{12},g_{13},g_{14},\dots,g_{n4}) where gi1,gi2g_{i1},g_{i2}, gi3g_{i3} and gi4g_{i4} are the four entries of the 2×22\times 2 matrix ρ(γi)\rho(\gamma_{i}). Fix a generating set for π1(M)\pi_{1}(\partial M), {M1,L1,,Mk,Lk}\{M_{1},L_{1},\dots,M_{k},L_{k}\} where each pair {Mi,Li}\{M_{i},L_{i}\} generates π1(Ti)\pi_{1}(T_{i}), the fundamental group a connected component of M\partial M. Let mi±1m_{i}^{\pm 1} be the eigenvalues of ρ(Mi)\rho(M_{i}) and li±1l_{i}^{\pm 1} be the eigenvalues of ρ(Li)\rho(L_{i}).

Define the following functions on (g11,,gn4,m1±1,l1±1,,mk±1,lk±1)\mathbb{C}(g_{11},\dots,g_{n4},m_{1}^{\pm 1},l_{1}^{\pm 1},\dots,m_{k}^{\pm 1},l_{k}^{\pm 1})

IMi(mi+mi1),ILi(li+li1),IMiLi(mili+mi1li1).I_{M_{i}}-(m_{i}+m_{i}^{-1}),\quad I_{L_{i}}-(l_{i}+l_{i}^{-1}),\quad I_{M_{i}L_{i}}-(m_{i}l_{i}+m_{i}^{-1}l_{i}^{-1}).

The extended representation variety, RE(M)R_{E}(M) is the algebraic set in 4n×()2k\mathbb{C}^{4n}\times(\mathbb{C}^{*})^{2k} cut out by 𝔘R(M)\mathfrak{U}_{R(M)} and the equations above for i=1,,ki=1,\dots,k. The extended character variety XE(M)X_{E}(M) is defined similarly. The eigenvalue variety, E(M)E(M), is the closure of the image of RE(M)R_{E}(M) under the projection to the coordinates m1,l1,,mk,lkm_{1},l_{1},\dots,m_{k},l_{k}. It naturally sits in ()2k(m1±1,l1±1,,mk±1,lk±1)(\mathbb{C}^{*})^{2k}(m_{1}^{\pm 1},l_{1}^{\pm 1},\dots,m_{k}^{\pm 1},l_{k}^{\pm 1}). We have the following commuting diagram where the maps p1p_{1}, p2p_{2}, and p3p_{3} are all finite to one of the same degree.

RE(M)\textstyle{R_{E}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}tE\scriptstyle{t_{E}}p1\scriptstyle{p_{1}}XE(M)\textstyle{X_{E}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iE\scriptstyle{i_{E}^{*}}p2\scriptstyle{p_{2}}E(M)\textstyle{E(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p3\scriptstyle{p_{3}}R(M)\textstyle{R(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}t\scriptstyle{t}X(M)\textstyle{X(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i^{*}}A(M)\textstyle{A(M)}

An irreducible component of R(M)R(M) containing a discrete and faithful representation is denoted R0(M)R_{0}(M), and the image in X(M)X(M) is denoted X0(M)X_{0}(M) and called a canonical component. For a given MM there can be more than one canonical component, as there are multiple lifts of a discrete faithful representation to SL2()\text{SL}_{2}(\mathbb{C}). We call an irreducible component of A(M)A(M) a canonical component and write A0(M)A_{0}(M) if it is the image of a canonical component in X(M)X(M). This definition also extends to the extended varieties.

We can define the PSL2()\text{PSL}_{2}(\mathbb{C}) representation and character varieties in a similar way. (See [2].) We denote the PSL2()\text{PSL}_{2}(\mathbb{C}) character variety of MM by Y(M)Y(M) and a canonical component by Y0(M)Y_{0}(M). We can form the restriction variety B(M)B(M) analogous to A(M)A(M) as above (and also form B0(M)B_{0}(M) when appropriate). The group PSL2()\text{PSL}_{2}(\mathbb{C}) is isomorphic to Isom+(3)\rm{Isom}^{+}(\mathbb{H}^{3}), the group of orientation preserving isometries of hyperbolic half space. Since the index [Isom(3):Isom+(3)]=2[\rm{Isom}(\mathbb{H}^{3}):\rm{Isom}^{+}(\mathbb{H}^{3})]=2, for MM a finite volume hyperbolic 3-manifold with at least one cusp there are two PSL2()\text{PSL}_{2}(\mathbb{C}) characters corresponding to discrete faithful representations of π1(M)\pi_{1}(M). These correspond to different orientations of MM. The authors know of no examples where these two characters lie on different irreducible components of Y(M)Y(M). The coordinate ring of the quotient is TeT_{e}\otimes\mathbb{C} where TeT_{e} is the subring of TT consisting of all elements invariant under the action of μ2\mu_{2}. Culler [6] (Corollary 2.3) showed that a faithful representation of any discrete torsion-free subgroup to PSL2()\text{PSL}_{2}(\mathbb{C}) lifts to SL2()\text{SL}_{2}(\mathbb{C}), and moreover (Theorem 4.1) that a representation ρ:ΓPSL2()\rho:\Gamma\rightarrow\text{PSL}_{2}(\mathbb{C}) lifts to a representation to SL2()\text{SL}_{2}(\mathbb{C}) if and only if every representation in the path component containing ρ\rho lifts.

By the proof of Thurston’s Dehn Surgery Theorem ([24] Theorem 5.8.2), any neighborhood of the character of a discrete faithful representation in Y(M)Y(M) contains all but finitely many Dehn filling characters. Since a discrete faithful character is always a simple point of Y(M)Y(M) [21] it follows Y0(M(,r))Y_{0}(M(-,r)) is contained in some Y0(M)Y_{0}(M) for all but finitely many rr. Therefore, for all but finitely many rr, X0(M(,r))X_{0}(M(-,r)) is contained in some canonical component of X(M)X(M). (Due to lifting restrictions it is not necessarily true that all but finitely many X0(M(,r))X_{0}(M(-,r)) are contained in a given X0(M)X_{0}(M).)

In the light of Lemma 3.5 we now establish bounds on the degrees of maps between the varieties of interest.

Lemma 4.2.

Let MM be a finite volume hyperbolic 3-manifold with exactly two cusps. Assume that M(,r)M(-,r) is hyperbolic.

  1. (1)

    Y0(M(,r)B0(M(,r))Y_{0}(M(-,r)\rightarrow B_{0}(M(-,r)) has degree 1.

  2. (2)

    X0(M(,r))A0(M(,r))X_{0}(M(-,r))\rightarrow A_{0}(M(-,r)) has degree at most 12|H1(M;/2)|\tfrac{1}{2}|H^{1}(M;\mathbb{Z}/2\mathbb{Z})|.

  3. (3)

    E0(M(,r))A0(M(,r))E_{0}(M(-,r))\rightarrow A_{0}(M(-,r)) has degree 4.

  4. (4)

    X0(M(,r))Y0(M(,r))X_{0}(M(-,r))\rightarrow Y_{0}(M(-,r)) has degree at most |H1(M;/2)||H^{1}(M;\mathbb{Z}/2\mathbb{Z})|.

These mappings are all dense.

Proof.

Part (1) follows directly from Dunfield [7] (Theorem 3.1), who showed that the mapping i:Y0(M(,r))B0(M(,r))i^{*}:Y_{0}(M(-,r))\rightarrow B_{0}(M(-,r)) is a birational map onto its image. Dunfield [7] (Corollary 3.2), also proved that the degree of the map i:X0(M(,r))A0(M(,r))i^{*}:X_{0}(M(-,r))\rightarrow A_{0}(M(-,r)) is at most 12|H1(M(,r);/2)|\tfrac{1}{2}|H^{1}(M(-,r);\mathbb{Z}/2\mathbb{Z})|. By Lemma 4.1 this is bounded by 12|H1(M;/2)|\tfrac{1}{2}|H^{1}(M;\mathbb{Z}/2\mathbb{Z})|. This establishes part (2). The mapping from E0(M(,r))E_{0}(M(-,r)) to X0(M(,r))X_{0}(M(-,r)) is given by the action of inverting both entries of the pair (m1,l1)(m_{1},l_{1}) to (m11,l11)(m_{1}^{-1},l_{1}^{-1}). This has degree at most four, giving part (3). Part (4) follows from the observation that the number of lifts of a representation ρ:ΓPSL2()\rho:\Gamma\rightarrow\text{PSL}_{2}(\mathbb{C}) to SL2()\text{SL}_{2}(\mathbb{C}) is |H1(Γ;/2)||H^{1}(\Gamma;\mathbb{Z}/2\mathbb{Z})|. Therefore the degree of the map is at most |H1(M(,r);/2)||H^{1}(M(-,r);\mathbb{Z}/2\mathbb{Z})| which is bounded above by |H1(M;/2)||H^{1}(M;\mathbb{Z}/2\mathbb{Z})| by Lemma 4.1. since the maps commute.

The mappings must all have dense image in the target space since all maps considered are maps between curves and the image must have dimension one as the degrees of the maps are bounded. ∎

Given Lemma 4.2, and Definition 3.6 we record the following for convenience.

Proposition 4.3.

Let MM be a finite volume hyperbolic 3-manifold with exactly two cusps. Assume that M(,r)M(-,r) is hyperbolic. Define the set

𝔖={A0(M(,r)),B0(M(,r)),E0(M(,r)),X0(M(,r)),Y0(M(,r))}.\mathfrak{S}=\{A_{0}(M(-,r)),B_{0}(M(-,r)),E_{0}(M(-,r)),X_{0}(M(-,r)),Y_{0}(M(-,r))\}.

For any V,W𝔖V,W\in\mathfrak{S}, γ(V)γ(W)\gamma(V)\sim\gamma(W).

Proof.

By Lemma 3.5 it suffices to show that for all but finitely many rr that there is a dominant map from VV to WW or from WW to VV whose degree is independent of rr. The lemma now follows from Lemma 4.2. ∎

5. Isolation of cusps

Let MM be a finite volume hyperbolic 3-manifold with kk cusps. Geometric isolation was studied extensively in [17].

Definition 5.1.

Cusps 1,,l1,\dots,l are geometrically isolated from cusps l+1,kl+1,\dots k if any deformation induced by Dehn filling cusps l+1,kl+1,\dots k while keeping cusps 1,l1,\dots l complete does not change the Euclidean structures at cusps 1,,l1,\dots,l.

Cusps 1,,l1,\dots,l are first order isolated from cusps l+1,,kl+1,\dots,k if the map from the space of deformations induced by Dehn filling cusps l+1,,kl+1,\dots,k while keeping cusps 1,,l1,\dots,l complete to the space of Euclidean structures at cusps 1,,l1,\dots,l has zero derivative at the point corresponding to the complete structure on MM.

Cusps 1,,l1,\dots,l are strongly geometrically isolated from cusps l+1,,kl+1,\dots,k if performing integral Dehn filling of cusps 1,,l1,\dots,l and replacing the cusps by goedesics γ1,,γl\gamma_{1},\dots,\gamma_{l} and then deforming cusps l+1,,kl+1,\dots,k does not change the geometry of γ1,,γl\gamma_{1},\dots,\gamma_{l}.

Strong geometric isolation implies geometric isolation which implies first order isolation. Strong geometric isolation and first order geometric isolation are symmetric conditions, but geometric isolation is not necessarily symmetric.

We now restrict to our case of interest, namely the case when MM has exactly two cusps, T1T_{1} and T2T_{2}. The projection ϖi(A0(M))\varpi_{i}(A_{0}(M)) is either a curve or is dense in 2\mathbb{C}^{2} (see Lemma 6.5 below). This projection is a curve exactly when the projection ϖi(B0(M))\varpi_{i}(B_{0}(M)) is a curve, and therefore this dimension is well-defined even if there are multiple A0(M)A_{0}(M) components.

We will adopt the notation from [17]. Define the set

S={(u1,τ1(u1,u2),u2,τ2(u1,u2))}4(u1,w1,u2,w2)S=\{(u_{1},\tau_{1}(u_{1},u_{2}),u_{2},\tau_{2}(u_{1},u_{2}))\}\subset\mathbb{C}^{4}(u_{1},w_{1},u_{2},w_{2})

where uiu_{i} is twice the logarithm of the eigenvalue of the holonomy of the meridian of TiT_{i} and viv_{i} is twice the logarithm of the eigenvalue of the holonomy of the longitude of TiT_{i}, and τi=ui/vi\tau_{i}=u_{i}/v_{i}.

Lemma 5.2.

The first cusp of MM is geometrically isolated from the second cusp of MM if and only if ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is a curve.

Proof.

The uiu_{i} coordinates correspond to the meridional mim_{i} coordinates. The τi\tau_{i} coordinates are of the form ui/viu_{i}/v_{i} and therefore (as long as uivi0u_{i}v_{i}\neq 0) are in bijective correspondence with the viv_{i} coordinates. These, in turn, give the longitudinal coordinates, lil_{i}. Therefore, the dimension of A0(M)A_{0}(M) is the same as the dimension of SS, and the dimension of ϖi(A0(M))\varpi_{i}(A_{0}(M)) equals the dimension of ϖi(S)\varpi_{i}(S).

The projection ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is {(u1,τ1(u1,u2))}\{(u_{1},\tau_{1}(u_{1},u_{2}))\}. This is a curve if and only if τ1\tau_{1} is a function of u1u_{1} alone, and not a function of u2u_{2}. In this case τ1(0,u2)\tau_{1}(0,u_{2}) is constant. This is the condition for T1T_{1} to be geometrically isolated from T2T_{2} (see [17], proof of Theorem 4.2).

Assume that T1T_{1} is geometrically isolated from T2T_{2}. The projection ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is a curve exactly when {(u1,τ1(u1,u2))}\{(u_{1},\tau_{1}(u_{1},u_{2}))\} is a curve. In this case, by [17] τ1(0,u2)\tau_{1}(0,u_{2}) is a constant, cc. Therefore the line {(0,w1)}2(u1,w1)\{(0,w_{1})\}\subset\mathbb{C}^{2}(u_{1},w_{1}) intersects {(u1,τ1(u1,u2))}\{(u_{1},\tau_{1}(u_{1},u_{2}))\} in a single point, (0,c)(0,c). It follows that {(u1,τ1(u1,u2))}\{(u_{1},\tau_{1}(u_{1},u_{2}))\} cannot be dense in 2\mathbb{C}^{2}. (A Riemannian neighborhood of any point in {(u1,τ(u1,u2))}\{(u_{1},\tau(u_{1},u_{2}))\} is homeomorphic to 2\mathbb{C}^{2} if the image is dense. However, by the above, a neighborhood of (0,c)(0,c) cannot be homeomorphic to 2\mathbb{C}^{2}.) Therefore, the image has dimension 1. As A0(M)A_{0}(M) is irreducible by construction, by Lemma 3.2 we conclude that ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is a curve as well.

Lemma 5.3.

The cusps of MM are strongly geometrically isolated from one another if and only if A0(M)C1×C2A_{0}(M)\cong C_{1}\times C_{2} where CiC_{i} is a curve in 2(mi,li)\mathbb{C}^{2}(m_{i},l_{i}).

Proof.

By [17] Theorem 4.3, T1T_{1} is strongly geometrically isolated from T2T_{2} if and only if v1v_{1} is dependent on u1u_{1} and not u2u_{2}. Strong geometric isolation is a symmetric condition (see [17]). Therefore, T1T_{1} is strongly geometrically isolated from T2T_{2} if and only if m1m_{1} depends only on l1l_{1} and m2m_{2} depends only on l2l_{2}. Therefore, for any irreducible polynomial ff in the vanishing ideal of A0(M)A_{0}(M), ff is a polynomial of m1m_{1} and l1l_{1} or of m2m_{2} and l2l_{2}. It follows that A0(M)A_{0}(M) is the product of curves if and only if T1T_{1} is strongly geometrically isolated from T2T_{2}.

Corollary 5.4.

The cusps of MM are strongly geometrically isolated from one another if and only if both cusps are geometrically isolated from one another.

Proof.

If the cusps of MM are strongly geometrically isolated, then by Lemma 5.3 A0(M)C1×C2A_{0}(M)\cong C_{1}\times C_{2} is a product of curves, so that ϖi(A0(M))=Ci\varpi_{i}(A_{0}(M))=C_{i} is a curve for i=1,2i=1,2 and by Lemma 5.2 both cusps are geometrically isolated from each other. If both cusps are geometrically isolated from each other, then ϖi(A0(M))\varpi_{i}(A_{0}(M)) is a curve, say CiC_{i} for i=1,2i=1,2. The surface A0(M)A_{0}(M) is contained in both ϖ1(C1)1=C1×\varpi_{1}(C_{1})^{-1}=C_{1}\times\mathbb{C} and ϖ2(C2)1=×C2\varpi_{2}(C_{2})^{-1}=\mathbb{C}\times C_{2}. The intersection of these sets is C1×C2C_{1}\times C_{2}, so A0(M)A_{0}(M) is a product of curves. By Lemma 5.3 we conclude that the cusps of MM are strongly geometrically isolated.

6. Gonality in Dehn surgery space

In this section we restrict to the case where MM is a finite volume hyperbolc 3-manifold with exactly two cusps. Before we proceed, we define some Chebyshev polynomials which will appear as traces.

Definition 6.1.

For any integer kk let fk(x)f_{k}(x) be the kthk^{th} Fibonacci polynomial, that is defined f0(x)=0f_{0}(x)=0, f1(x)=1f_{1}(x)=1 and for all other kk, define fk(x)f_{k}(x) recursively by the relation

fk+1(x)+fk1(x)=xfk(x).f_{k+1}(x)+f_{k-1}(x)=xf_{k}(x).

Define gk=fkfk1g_{k}=f_{k}-f_{k-1}, and hk(x)=fk+1fk1h_{k}(x)=f_{k+1}-f_{k-1}.

Remark 6.2.

With x=s+s1x=s+s^{-1} then

fk(x)=skskss1,gk(x)=sk+s1ks+1,andhk(x)=sk+sk.f_{k}(x)=\frac{s^{k}-s^{-k}}{s-s^{-1}},\quad g_{k}(x)=\frac{s^{k}+s^{1-k}}{s+1},\quad\text{and}\quad h_{k}(x)=s^{k}+s^{-k}.

The effect of r=p/qr=p/q Dehn filling of the second cusp of MM on the fundamental group is that it introduces the group relation M2pL2q=1M_{2}^{p}L_{2}^{q}=1. Therefore, for a representation ρ\rho the relation introduces an additional matrix relation ρ(M2)pρ(L2)q=I2\rho(M_{2})^{p}\rho(L_{2})^{q}=I_{2} where I2I_{2} is the 2×22\times 2 identity matrix. Up to conjugation, ρ(M2)\rho(M_{2}) can be taken to be upper triangular, and since M2M_{2} and L2L_{2} commute, ρ(L2)\rho(L_{2}) is upper triangular as well. (If ρ(M2)=I2\rho(M_{2})=I_{2} we can conjugate so that ρ(L2)\rho(L_{2}) is upper triangular.) The character and eigenvalue varieties are not changed by conjugation. Therefore, we may take

ρ(M2)=(ms0m1)andρ(L2)=(lt0l1)\rho(M_{2})=\left(\begin{array}[]{cc}m&s\\ 0&m^{-1}\end{array}\right)\quad\text{and}\quad\rho(L_{2})=\left(\begin{array}[]{cc}l&t\\ 0&l^{-1}\end{array}\right)

where m,lm,l\in\mathbb{C}^{*} and s,ts,t\in\mathbb{C}. The commuting condition ensures that

(1) t(mm1)=s(ll1).t(m-m^{-1})=s(l-l^{-1}).

By the Cayley-Hamilton theorem,

ρ(M2)p=(mpsfp(m+m1)0mp)andρ(L2)q=(lqtfq(l+l1)0lq).\rho(M_{2})^{p}=\left(\begin{array}[]{cc}m^{p}&sf_{p}(m+m^{-1})\\ 0&m^{-p}\end{array}\right)\quad\text{and}\quad\rho(L_{2})^{-q}=\left(\begin{array}[]{cc}l^{-q}&-tf_{q}(l+l^{-1})\\ 0&l^{q}\end{array}\right).

As ρ(M2)p=ρ(M2)q\rho(M_{2})^{p}=\rho(M_{2})^{-q}, we conclude that

mplq=1 and sfp(m+m1)=tfq(l+l1).m^{p}l^{q}=1\ \text{ and }\ sf_{p}(m+m^{-1})=-tf_{q}(l+l^{-1}).

If neither mm nor ll is ±1\pm 1 then (1) implies that

smpmpmm1=tmpmpll1.s\frac{m^{p}-m^{-p}}{m-m^{-1}}=t\frac{m^{p}-m^{-p}}{l-l^{-1}}.

With mplq=1m^{p}l^{q}=1 the condition sfp(m+m1)=tfq(l+l1)sf_{p}(m+m^{-1})=-tf_{q}(l+l^{-1}) follows. Therefore, the relation sfp(m+m1)=tfq(l+l1)sf_{p}(m+m^{-1})=-tf_{q}(l+l^{-1}) follows from (1) and mplq=1m^{p}l^{q}=1 unless mm or ll is ±1\pm 1. If m=±1m=\pm 1 or l=±1l=\pm 1, an additional identity with ss and tt is required to deduce sfp(m+m1)=tfq(l+l1)sf_{p}(m+m^{-1})=-tf_{q}(l+l^{-1}) from the commuting condition and mplq=1m^{p}l^{q}=1. (The commuting condition (1) is part of the defining equations for X(M)X(M) as MiM_{i} and LiL_{i} commute in the group.) In the next section, we will see how this additional relation manifests as an intersection in the character varieties. First, we will discuss the containment of these sets.

The set X(M)X(\partial M) is contained in 4(m1,l1,m2,l2)\mathbb{C}^{4}(m_{1},l_{1},m_{2},l_{2}) where the coordinates are distinguished eigenvalues of the meridians and longitudes of the cusps. Any irreducible component has dimension at most two, and the image of a canonical component has dimension two (See [23] and [25] Proposition 12). The set A(M(,r))A(M(-,r)) is contained in 2(m,l)\mathbb{C}^{2}(m,l) where mm and ll are identified with m1m_{1} and l1l_{1}, and any canonical component has dimension one.

For all but finitely many r,r, Y0(M(,r))Y_{0}(M(-,r)) is contained in some Y0(M)Y_{0}(M). Therefore for all but finitely many rr, X0(M(,r))X_{0}(M(-,r)) is contained in some X0(M)X_{0}(M). There may be numerous lifts of Y0(M)Y_{0}(M), each containing infinitely many X0(M(,r))X_{0}(M(-,r)). As A0(M)4A_{0}(M)\subset\mathbb{C}^{4} and A0(M(,r))2A_{0}(M(-,r))\subset\mathbb{C}^{2} it is not the case that A0(M(,r))A0(M)A_{0}(M(-,r))\subset A_{0}(M). In Lemma 6.5 we will show the relationship between these two sets.

6.1. Filling Relations

Recall the notation established in §3; if ff is a polynomial, V(f)V(f) indicates the vanishing set of (the ideal generated by) ff in a specified n\mathbb{C}^{n}. We will make extensive use of the sets and polynomials from Definition 3.7. We will write V(m2±pl2q1)V(m_{2}^{\pm p}l_{2}^{q}-1) for V(m2|p|l2q1)V(m2|p|l2q)V(m_{2}^{|p|}l_{2}^{q}-1)\cup V(m_{2}^{|p|}-l_{2}^{q}).

Lemma 6.3.

Assume that M(,r)M(-,r) is hyperbolic. For all but finitely many rr there is some X0(M)X_{0}(M) such that

X0(M(,r))X0(M)(V(m2±pl2q1)).X_{0}(M(-,r))\subset X_{0}(M)\cap\Big{(}V(m_{2}^{\pm p}l_{2}^{q}-1)\Big{)}.
Proof.

For all but perhaps finitely many rr such that M(,r)M(-,r) is hyperbolic there is some X0(M)X_{0}(M) containing a given X0(M(,r))X_{0}(M(-,r)). Consider these r=p/qr=p/q. The fundamental group π1(M(,r))\pi_{1}(M(-,r)) is the quotient of π1(M)\pi_{1}(M) corresponding to the addition of a single relation, M2pL2q=1M_{2}^{p}L_{2}^{q}=1. Therefore, the trace of ρ(M2)p\rho(M_{2})^{p} equals the trace of ρ(L2)q\rho(L_{2})^{-q}. By construction, these traces are hp(χρ(M2))h_{p}(\chi_{\rho}(M_{2})) and hq(χρ(L2))h_{q}(\chi_{\rho}(L_{2})) with χρ(M2)=m2+m21\chi_{\rho}(M_{2})=m_{2}+m_{2}^{-1} and χρ(L2)=l2+l21\chi_{\rho}(L_{2})=l_{2}+l_{2}^{-1}. This shows that

X0(M(,r))X0(M)V(hp(m2+m21)hq(l2+l21)).X_{0}(M(-,r))\subset X_{0}(M)\cap V\big{(}h_{p}(m_{2}+m_{2}^{-1})-h_{q}(l_{2}+l_{2}^{-1})\big{)}.

It suffices to show that

V(hp(m2+m21)hq(l2+l21))V(m2±pl2q1).V\big{(}h_{p}(m_{2}+m_{2}^{-1})-h_{q}(l_{2}+l_{2}^{-1})\big{)}\subset V(m_{2}^{\pm p}l_{2}^{q}-1).

This is equivalent to showing that the solution set of m2p+m2p=l2q+l2qm_{2}^{p}+m_{2}^{-p}=l_{2}^{q}+l_{2}^{-q} is contained in the solution set of m2±pl2q=1m_{2}^{\pm p}l_{2}^{q}=1 in 4(m2,l2)\mathbb{C}^{4}(m_{2},l_{2}). This follows directly from the observation that the solutions to an equation of the form x+x1=y+y1x+x^{-1}=y+y^{-1} are x=y±1x=y^{\pm 1}.

Remark 6.4.

It is possible that other trace relations will specify the sign in the relation m2±pl2q1m_{2}^{\pm p}l_{2}^{q}-1 but this is not necessarily the case.

We will assume that MM is hyperbolic and X0(M(,r))X0(M)X_{0}(M(-,r))\subset X_{0}(M). The sets A0(M(,r))A_{0}(M(-,r)) and A0(M)A_{0}(M) lie in different ambient spaces, but we have the following lemma which allows us to concretely relate A0(M(,r))A_{0}(M(-,r)) with A(M)A(M). Since MM is hyperbolic A0(M)A_{0}(M) is a surface, and for all but finitely many rr, A0(M(,r))A_{0}(M(-,r)) is a curve. As a result, the projection map ϖi(A0(M))\varpi_{i}(A_{0}(M)) has dimension at most two. The following lemma will allow us to deduce that ϖi(A0(M))\varpi_{i}(A_{0}(M)) cannot be a point, and therefore is either a curve or is dense in 2\mathbb{C}^{2}. (The image is irreducible by Lemma 3.2.) Our analysis will depend upon the dimension of these images.

Lemma 6.5.

Assume that M(,r)M(-,r) is hyperbolic. For all but finitely many rr, A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) is a finite union of curves and the variety A0(M(,r))A_{0}(M(-,r)) is a curve component of the closure of ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)).

Proof.

The group π1(M(,r))\pi_{1}(M(-,r)) is a quotient of π1(M)\pi_{1}(M); the group presentation only has the additional relation that M2pL2q=1M_{2}^{p}L_{2}^{q}=1, where r=p/qr=p/q. Using the notation established earlier in this section, in the construction of X(M)X(\partial M) and A(M(,r))A(M(-,r)) the (1,1)(1,1) entry of ρ(Mi)\rho(M_{i}) is identified with mim_{i}, and the (1,1)(1,1) entry for ρ(Li)\rho(L_{i}) is identified with lil_{i}. The manifestation of this group quotient on the character varieties is that X0(M(,r))X_{0}(M(-,r)) is contained in the intersection of X0(M)X_{0}(M) and V(m2±pl2q1)V(m_{2}^{\pm p}l_{2}^{q}-1), as shown in Lemma 6.3.

The epimorphism π1(M)π1(M(,r))\pi_{1}(M)\rightarrow\pi_{1}(M(-,r)) surjects π1(M)\pi_{1}(\partial M) onto π1(M(,r))\pi_{1}(\partial M(-,r)) and this induces maps

X(M)iX(M)ϖ1X(M(,r)).X(M)\overset{i^{*}}{\rightarrow}X(\partial M)\overset{\varpi_{1}}{\rightarrow}X(\partial M(-,r)).

The set X(M)X(\partial M) is dense in 4(m1,l1,m2,l2)\mathbb{C}^{4}(m_{1},l_{1},m_{2},l_{2}) and the set X(M(,r))X(\partial M(-,r)) is contained in 2(m1,l1)\mathbb{C}^{2}(m_{1},l_{1}).

Since X0(M(,r))X0(M)V(m2±pl2q1)X_{0}(M(-,r))\subset X_{0}(M)\cap V(m_{2}^{\pm p}l_{2}^{q}-1), it follows that

i(X0(M(,r)))i(X0(M))i(V(m2±pl2q1))).i_{*}(X_{0}(M(-,r)))\subset i_{*}(X_{0}(M))\cap i_{*}(V(m_{2}^{\pm p}l_{2}^{q}-1))).

Therefore, i(X0(M(,r)))A0(M)V(m2±pl2q1)i_{*}(X_{0}(M(-,r)))\subset A_{0}(M)\cap V(m_{2}^{\pm p}l_{2}^{q}-1) in 4\mathbb{C}^{4}. Finally, since A0(M(,r))A_{0}(M(-,r)) is the closure of ϖ1(i(X0(M(,r))))\varpi_{1}(i_{*}(X_{0}(M(-,r)))), by projecting we conclude that A0(M(,r))A_{0}(M(-,r)) is contained in (the closure of) ϖ1(A0(M)V(m2±pl2q1))\varpi_{1}(A_{0}(M)\cap V(m_{2}^{\pm p}l_{2}^{q}-1)), which is ϖ1(A0(M)(r)).\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)).

To show that A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) is a finite union of curves, by Lemma 3.11 it suffices to show dim(ϖi(A0(M)))>0\dim_{\mathbb{C}}(\varpi_{i}(A_{0}(M)))>0 for i=1,2i=1,2. Since M(,r)M(-,r) is hyperbolic and has exactly one cusp A0(M(,r))A_{0}(M(-,r)) is a curve. From the above,

A0(M(,r))ϖ1(A0(M)(r))ϖ1(A0(M))A_{0}(M(-,r))\subset\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r))\subset\varpi_{1}(A_{0}(M))

which implies that dimϖ1(A0(M))>0\dim_{\mathbb{C}}\varpi_{1}(A_{0}(M))>0. By symmetry (filling the other cusp), dimϖ2(A0(M))>0\dim_{\mathbb{C}}\varpi_{2}(A_{0}(M))>0 as well.

Refer to caption
Figure 1. The set A0(M)4(m1,l1,m2,l2)A_{0}(M)\subset\mathbb{C}^{4}(m_{1},l_{1},m_{2},l_{2}) with the intersection A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) and the projection ϖ1(A0(M))2(m1,l1)\varpi_{1}(A_{0}(M))\subset\mathbb{C}^{2}(m_{1},l_{1}) when the first cusp is geometrically isolated from the second. The set A0(M(r))ϖ1(A0(M))A_{0}(M(r))\subset\varpi_{1}(A_{0}(M)).

The following lemma shows that if the first cusp of MM is geometrically isolated from the second cusp of MM then there are only finitely many A0(M(,r))A_{0}(M(-,r)). We conclude that in this situation, the gonality of any A0(M(,r))A_{0}(M(-,r)) is bounded independently of rr.

Lemma 6.6.

Assume that the first cusp of MM is geometrically isolated from second cusp of MM. For all but finitely many rr such that A0(M(,r))ϖ1(A0(M))A_{0}(M(-,r))\subset\varpi_{1}(A_{0}(M)), all of the curves A0(M(,r))A_{0}(M(-,r)) are identical.

Proof.

Lemma 6.5 implies that for all but finitely many rr, A0(M(,r))A_{0}(M(-,r)) is contained in ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)). For all such rr, A0(M(,r))A_{0}(M(-,r)) is a curve and so is ϖ1(A0(M))\varpi_{1}(A_{0}(M)) by Lemma 5.2. It follows that they must be the same curve. ∎

Remark 6.7.

A similar statement applies to B0(M)B_{0}(M). If there is a single canonical component Y0(M)Y_{0}(M) then Lemma 6.6 shows that for all but finitely many rr the curves B0(M(,r))B_{0}(M(-,r)) are the same curve.

6.2. Gonality Relations

Now we show that if the first cusp of MM is not geometrically isolated from the second cusp of MM, in order to bound the gonality of A0(M(,r))A_{0}(M(-,r)) it suffices to bound the gonality of A0(M)(r)A_{0}(M)\cap\mathcal{H}(r).

Refer to caption
Figure 2. The set A0(M)4(m1,l1,m2,l2)A_{0}(M)\subset\mathbb{C}^{4}(m_{1},l_{1},m_{2},l_{2}) with the intersection A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) and the projection ϖ1(A0(M))2(m1,l1)\varpi_{1}(A_{0}(M))\subset\mathbb{C}^{2}(m_{1},l_{1}) when the first cusp is not geometrically isolated from the second. The set A0(M(r))ϖ1(A0(M))A_{0}(M(r))\subset\varpi_{1}(A_{0}(M)).
Lemma 6.8.

Assume that the first cusp of MM is not geometrically isolated from the second cusp of MM. Then for all but finitely many rr such that A0(M(,r))ϖ1(A0(M))A_{0}(M(-,r))\subset\varpi_{1}(A_{0}(M)) the set A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) is a finite union of curves. Additionally, for any irreducible curve component CC of A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) with A0(M(,r))A_{0}(M(-,r)) contained in the closure of ϖ1(C)\varpi_{1}(C)

γ(C)γ(A0(M(,r))).\gamma(C)\sim\gamma(A_{0}(M(-,r))).
Proof.

It suffices to prove the lemma for a single irreducible component A0(M)A_{0}(M) whose projection contains infinitely many A0(M(,r))A_{0}(M(-,r)). By Lemma 3.5 it suffices to show that with the exception of finitely many rr, the degree of ϖ1\varpi_{1} restricted to A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) is independent of rr.

By Lemma 5.2 since the first cusp of MM is not geometrically isolated from the second cusp of MM, the map ϖ1:A0(M)2\varpi_{1}:A_{0}(M)\rightarrow\mathbb{C}^{2} is dense. By Proposition 3.1 this is a dense map between surfaces of degree dd. Moreover, for all but finitely many points in the image of ϖ1(A0(M))\varpi_{1}(A_{0}(M)) the pre image of a point consists of at most dd points in A0(M)A_{0}(M). The image ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is 2\mathbb{C}^{2} with the possible exception of finitely many curves and points. By Lemma 6.5 except for finitely many points, A0(M(,r))A_{0}(M(-,r)) is contained in the image of ϖ1(A0(M))\varpi_{1}(A_{0}(M)).

Therefore, for all but finitely many points PP on any curve A0(M(,r))A_{0}(M(-,r)), the number of pre-images of PP under ϖ1\varpi_{1} is at most dd. As a result, if C(A0(M)(r))C\subset\big{(}A_{0}(M)\cap\mathcal{H}(r)\big{)} is an irreducible component such that A0(M(,r))A_{0}(M(-,r)) is contained in the closure of ϖ1(C)\varpi_{1}(C), the map ϖ1C\varpi_{1}\mid_{C} is degree at most dd. Since dd independent of rr, we have γ(C)γ(A0(M(,r)))\gamma(C)\sim\gamma(A_{0}(M(-,r))).

6.3. Proof of Theorem 1.1

Proof of Theorem 1.1.

To prove Theorem  1.1 it follows from Proposition 4.3 that it suffices to bound the gonality of A0(M(,r))A_{0}(M(-,r)). For all but finitely many r,r, X0(M(,r))X_{0}(M(-,r)) is contained in some X0(M)X_{0}(M). Therefore, for all but finitely many r,r, A0(M(,r))A_{0}(M(-,r)) is contained in some ϖ1(A0(M))\varpi_{1}(A_{0}(M)). As there are only finitely many A0(M)A_{0}(M), it suffices to consider all rr such that A0(M(,r))A_{0}(M(-,r)) is contained in a fixed ϖ1(A0(M))\varpi_{1}(A_{0}(M)).

Assume that neither cusp of MM is geometrically isolated from the other. By Lemma 5.2, A0(M)A_{0}(M) is a surface in 4\mathbb{C}^{4} whose image under both ϖ1\varpi_{1} and ϖ2\varpi_{2} is dense in 2\mathbb{C}^{2}. By Lemma 6.8 A0(M(,r))A_{0}(M(-,r)) is contained in the closure of ϖ1(C)\varpi_{1}(C) for some irreducible curve component CC of A0(M)(r)A_{0}(M)\cap\mathcal{H}(r), and γ(C)γ(A0(M(,r))\gamma(C)\sim\gamma(A_{0}(M(-,r)). By Lemma 3.13 since CA0(M)(r)C\subset A_{0}(M)\cap\mathcal{H}(r) and ϖ2(A0(M))\varpi_{2}(A_{0}(M)) is dense in 2\mathbb{C}^{2} it follows that γ(C)1\gamma(C)\sim 1. By the transitivity of \sim we conclude that γ(A0(M(,r)))1\gamma(A_{0}(M(-,r)))\sim 1.

Now we consider the case where there is one cusp geometrically isolated from another; one of the projections ϖi(A0(M))\varpi_{i}(A_{0}(M)) is a curve by Lemma 5.2.

Assume that ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is a curve, DD, and A0(M(,r))ϖ1(A0(M))A_{0}(M(-,r))\subset\varpi_{1}(A_{0}(M)). That is assume the first cusp is geometrically isolated from the second cusp. By Lemma 6.5, for all but finitely many rr, A0(M(,r))A_{0}(M(-,r)) is a curve component of the closure of ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)). We conclude that for all but finitely many of the rr above, ϖ1(A0(M)(r))=D\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r))=D as well, since this projection is a union of curves contained in DD. (In this case the map ϖ1A0(M)\varpi_{1}\mid_{A_{0}(M)} does not have finite degree and therefore we cannot conclude that γ(A0(M(,r)))γ(A0(M)(r))\gamma(A_{0}(M(-,r)))\sim\gamma(A_{0}(M)\cap\mathcal{H}(r)).) We conclude that γ(A0(M(,r))\gamma(A_{0}(M(-,r)) is bounded independent of rr since there are only finitely many such curves.

Now, assume that ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is dense in 2\mathbb{C}^{2} and the closure of ϖ2(A0(M))\varpi_{2}(A_{0}(M)) is a curve. By Lemma 5.2, this occurs when the second cusp is geometrically isolated from the first cusp, but the first cusp is not geometrically isolated from the second cusp. By Lemma 6.8, for any irreducible component CC of A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) with A0(M(,r))A_{0}(M(-,r)) contained in the closure of ϖ1(C)\varpi_{1}(C),

γ(C)γ(A0(M(,r))).\gamma(C)\sim\gamma(A_{0}(M(-,r))).

Therefore, it suffices to bound γ(C)\gamma(C). By Lemma 3.13, γ(C)1\gamma(C)\sim 1. (In this case the map ϖ2A0(M)\varpi_{2}\mid_{A_{0}(M)} does not have finite degree.) ∎

Remark 6.9.

The proof of Theorem 1.1 implies that if the first cusp is geometrically isolated from the second, then ϖ1(A0(M))\varpi_{1}(A_{0}(M)) is an irreducible curve and the curve A0(M(,r))A_{0}(M(-,r)) is the same for all rr such that ϖ1(A0(M))\varpi_{1}(A_{0}(M)) contains A0(M(,r))A_{0}(M(-,r)). A similar statement is true for B0(M(,r))B_{0}(M(-,r)). As there are only two characters of discrete faithful representations in Y0(M)Y_{0}(M), there are at most two B0(M)B_{0}(M) and at most two projections.

It is possible that all A0(M(,r))A_{0}(M(-,r)) are the same if the second cusp is geometrically isolated from the first cusp. This will occur if ϖ2(A0(M)(r))\varpi_{2}(A_{0}(M)\cap\mathcal{H}(r)) contains one of the common intersection points of the 𝒞(r)=ϖ2((r))\mathcal{C}(r)=\varpi_{2}(\mathcal{H}(r)). If not, we will show that the degree of ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)) is bounded by a constant because these curves are specializations of A0(M)A_{0}(M) at particular (m2,l2)(m_{2},l_{2}) values.

Remark 6.10.

By Proposition 4.3 the bounds in Theorem 1.1 translate to bounds for other sets as well. We obtain an analogous statement for each of the sets

A0(M(,r)),B0(M(,r)),E0(M(,r)),X0(M(,r)),Y0(M(,r)).A_{0}(M(-,r)),B_{0}(M(-,r)),E_{0}(M(-,r)),X_{0}(M(-,r)),Y_{0}(M(-,r)).

For a given rr such that M(,r)M(-,r) is hyperbolic, the correspondence between two such sets is established by a mapping between the sets. Even though this is ambiguous, any lift can be traced back to Y0Y_{0} so there is a correspondence between different lifts.

Remark 6.11.

Assume that MM has two cusps neither of which is geometrically isolated from the other. (We follow [4] and use the notation established above.) With r=p/qr=p/q, the core curve of the second cusp of M(,r)M(-,r) is given by γ=μ2bλ2a\gamma=\mu_{2}^{-b}\lambda_{2}^{a} with ap+qb=1ap+qb=1. We choose a parameter TT so that Tq=m2T^{-q}=m_{2} and Tp=l2T^{p}=l_{2}, so that m2pl2q=1m_{2}^{p}l_{2}^{q}=1. With ξ\xi the (distinguished) eigenvalue of ρ(γ)\rho(\gamma),

ξ=m2bl2a=Tbq+ap=T.\xi=m_{2}^{-b}l_{2}^{a}=T^{bq+ap}=T.

The map used to bound the gonality is

(m1,l1,m2,l2)(m2,l2).(m_{1},l_{1},m_{2},l_{2})\mapsto(m_{2},l_{2}).

The image under this mapping has gonality equal to one as seen by the mapping

(m2,l2)(m2bl2a,m2pl2q).(m_{2},l_{2})\mapsto(m_{2}^{b}l_{2}^{-a},m_{2}^{p}l_{2}^{q}).

Using the parametrization by TT this is

(m2,l2)=(Tq,Tp)(T1,1).(m_{2},l_{2})=(T^{-q},T^{p})\mapsto(T^{-1},1).

After a final projection onto the first coordinate, since T=ξT=\xi, we see that the map is given by (m1,l1,m2,l2)ξ1.(m_{1},l_{1},m_{2},l_{2})\mapsto\xi^{-1}. Since sending a single variable to its inverse is a birational map, this is birationally equivalent to sending (m1,l1,m2,l2)(m_{1},l_{1},m_{2},l_{2}) to ξ\xi.

7. Genus and Degree

The gonality of a curve is intimately related to both the (geometric) genus of the curve, and the degree of the curve. Like gonality, the genus is a birational invariant. The degree of a curve depends on the embedding of the curve. Assume that CC is a curve with a specified embedding into some ambient space. Let gg, dd, and γ\gamma be the genus, degree, and gonality of CC, respectively. The Brill-Noether theorem relates the genus and the gonality of a curve. It states that

γg+32.\gamma\leq\lfloor\frac{g+3}{2}\rfloor.

The genus degree formula (the adjunction formula) states that if CC is a smooth plane curve then

g=12(d1)(d2).g=\tfrac{1}{2}(d-1)(d-2).

Each singularity of order ss reduces the right hand side by 12s(s1)\tfrac{1}{2}s(s-1).

With Theorem 1.1 we can establish bounds for the genus. We have shown that the gonality is bounded. We now consider the variety A0(M(,r))A_{0}(M(-,r)) which naturally is contained in 2(m,l)\mathbb{C}^{2}(m,l).

Theorem 1.4.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, then there is a positive constant c1c_{1} depending only on MM and the framing of the second cusp such that

d(A0(M(,r)))c1h(r).d(A_{0}(M(-,r)))\leq c_{1}\cdot h(r).

If one cusp is geometrically isolated from the other cusp, then there is a positive constant c2c_{2} depending only on MM and the framing of the second cusp, such that

d(A0(M(,r)))c2.d(A_{0}(M(-,r)))\leq c_{2}.
Proof.

First, we will bound the degree of (any irreducible component of) the intersection of a surface XX in 4(x1,y1,x2,y2)\mathbb{C}^{4}(x_{1},y_{1},x_{2},y_{2}) with (r)\mathcal{H}(r). As the dimension of (r)\mathcal{H}(r) is three and the dimension of XX is two, the condition that they intersect in a curve implies that the intersection is proper. It is a consequence of Bézout’s theorem (see [9] Theorem I,7.7) that with r=p/qr=p/q, since deg(r)=max{|p|,|q|}\deg\mathcal{H}(r)=\max\{|p|,|q|\}, the degree of the intersection is at most

deg(X(r))degXdeg(r)=degXmax{|p|,|q|}.\deg(X\cap\mathcal{H}(r))\leq\deg X\cdot\deg\mathcal{H}(r)=\deg X\cdot\max\{|p|,|q|\}.

For any variety W4W\subset\mathbb{C}^{4} the degree of ϖ1(W)\varpi_{1}(W) is at most the degree of WW. To see this, assume that WW intersects a complimentary hyperplane HH in the point PP. Then ϖ1(W)\varpi_{1}(W) intersects ϖ1(H)\varpi_{1}(H) in ϖ1(P)\varpi_{1}(P). It suffices to see that we can take HH so that ϖ1(H)\varpi_{1}(H) is a complimentary hyperplane to ϖ1(W)\varpi_{1}(W) and that this respects multiplicities.

Therefore, with X=A0(M)X=A_{0}(M) by Lemma 6.5, A0(M(,r))A_{0}(M(-,r)) is contained in the closure of ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)). We conclude that

degA0(M(,r))deg(ϖ1(A0(M)(r)))deg(A0(M)(r))\deg A_{0}(M(-,r))\leq\deg(\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)))\leq\deg(A_{0}(M)\cap\mathcal{H}(r))

and by Bézout’s theorem

deg(A0(M)(r))degA0(M)max{|p|,|q|}.\deg(A_{0}(M)\cap\mathcal{H}(r))\leq\deg A_{0}(M)\max\{|p|,|q|\}.

This implies the first assertion.

Now assume that the second cusp is geometrically isolated from the first cusp. Therefore ϖ2(A0(M))\varpi_{2}(A_{0}(M)) is a curve. Let {φi}\{\varphi_{i}\} be a generating set for 𝔘A0(M)\mathfrak{U}_{A_{0}(M)} where φi=φi(x1,y1,x2,y2)\varphi_{i}=\varphi_{i}(x_{1},y_{1},x_{2},y_{2}). Then 𝔘A0(M)(r)\mathfrak{U}_{A_{0}(M)\cap\mathcal{H}(r)} is generated by {φi(x1,y1,ar,br)}\{\varphi_{i}(x_{1},y_{1},a_{r},b_{r})\} as in the proof of Lemma 3.13. The degree of A0(M)(r)A_{0}(M)\cap\mathcal{H}(r) is bounded by the degrees of these equations. The map ϖ1\varpi_{1} is a projection. It follows that degree of the the image ϖ1(A0(M)(r))\varpi_{1}(A_{0}(M)\cap\mathcal{H}(r)) is bounded by the degree (in x1x_{1} and y1y_{1}) of the ϖi(x1,y1,x2,y2)\varpi_{i}(x_{1},y_{1},x_{2},y_{2}).

If the first cusp is geometrically isolated from the second cusp, the result follows from Lemma 6.6. ∎

We can bound the genus using the genus-degree formula. Specifically, we have the following.

Theorem 1.2.

Let MM be a finite volume hyperbolic 3-manifold with two cusps. If M(,r)M(-,r) is hyperbolic, there is a positive constant cc depending only on MM and the framing of the second cusp such that

g(A0(M(,r)))ch(r)2.g(A_{0}(M(-,r)))\leq c\cdot h(r)^{2}.
Proof.

It suffices to consider curves of degree more than two, since the genus of plane curves of smaller degree at most two is bounded. The theorem follows directly from the genus-degree formula which implies that if CC is a plane curve of degree dd then the genus of CC is at most 12(d1)(d2)\tfrac{1}{2}(d-1)(d-2), which is less than 12d2\tfrac{1}{2}d^{2}. By Theorem 1.4 there is a constant cc such that dch(r)d\leq ch(r) so that the genus us bounded by 12c2h(r)2\tfrac{1}{2}c^{2}h(r)^{2}.

Remark 7.1.

For r=p/qr=p/q with pq0pq\neq 0 this is

g(A0(M(,r)))cmax{p2,q2}.g(A_{0}(M(-,r)))\leq c\max\{p^{2},q^{2}\}.

This statement can also be made for B0(M(,r))B_{0}(M(-,r)). These genus bounds do not appear sharp. In § 9.2 we will show that the double twist knots the genus is linear in the number of twists in each twist region. Similarly, the genus bounds from [1] are linear in h(r)h(r).

8. Newton Polygons

Bounds for the genus and the gonality of a plane curve can be obtained from the associated Newton Polygon. Let f[x±1,y±1]f\in\mathbb{C}[x^{\pm 1},y^{\pm 1}] be an irreducible Laurent polynomial which defines a curve V(f)V(f) in ({0})2(\mathbb{C}-\{0\})^{2}. Let Δ(f)\Delta(f) be the associated Newton polygon of ff. In 1893, Baker related the genus of such a curve to the volume of the Newton Polygon.

Theorem 8.1 (Baker).

The geometric genus of V(f)V(f) is at most the number of lattice points in the interior of Δ(f)\Delta(f).

Recently, Castryck and Cools showed a similar relationship for the gonality of V(f)V(f). A \mathbb{Z}-affine transformation is a map from 2\mathbb{R}^{2} to 2\mathbb{R}^{2} of the form xAx+bx\mapsto Ax+b where AGL2()A\in\text{GL}_{2}(\mathbb{Z}) and b2b\in\mathbb{Z}^{2}. We say that two lattice polygons Δ\Delta and Δ\Delta^{\prime} are equivalent if there is a \mathbb{Z}-affine transformation φ\varphi such that φ(Δ)=Δ\varphi(\Delta)=\Delta^{\prime}. The lattice width of Δ\Delta is the smallest integer s0s\geq 0 such that there is a \mathbb{Z}-affine transformation φ\varphi such that φ(Δ)\varphi(\Delta) is contained in the horizontal strip {(x,y)2:0ys}\{(x,y)\in\mathbb{R}^{2}:0\leq y\leq s\}.

Theorem 8.2 (Castryck-Cools).

The gonality of V(f)V(f) is at most the lattice width of Δ(f)\Delta(f).

Culler and Shalen [5] showed that if MM has only one cusp, then X0(M)X_{0}(M) determines a norm on H1(M,)H_{1}(\partial M,\mathbb{R}). This norm encodes many topological properties of MM including information about the Dehn fillings of MM. This construction can be applied to other curve components of X(M)X(M) yielding a semi norm in general. Culler and Shalen, and Boyer and Zhang [2] used these semi norms to study surfaces in 3-manifolds. Boyer and Zhang also extended this semi norm to Y(M)Y(M).

We observe the following about non-norm curve components.

Theorem 8.3.

Let MM be a compact, connected, irreducible, \partial-irreducible 3-manifold with boundary torus. Let YY(M)Y\subset Y(M) be a curve which is not a norm curve. Then YY is birational to \mathbb{C}.

Proof.

By Boyer and Zhang [2] (Proposition 5.4 3), if MM has one cusp any norm curve of X(M)X(M) maps to a component of the AA-polynomial whose Newton polygon is two-dimensional, and any curve that is not a norm curve maps to a component of the AA-polynomial that is one-dimensional or zero-dimensional. Therefore, the Newton polygon of a non-norm curve is a line segment or a point. This has empty interior, so by Baker’s theorem, or Castryck and Cools’ Theorem, the geometric genus is zero. It follows that the curve is birational to \mathbb{C}. ∎

9. Examples

Theorem 1.1 shows that the gonality of X0(M(,r))X_{0}(M(-,r)) is bounded independent of rr. In general, it is difficult to determine the exact gonality of such curves, as explicit computations of character varieties are difficult and only a few infinite families have been computed. In this section, we will compute the gonality of the character varieties associated to two families of manifolds which arise as Dehn surgeries. These examples will demonstrate that the linear bounds in Theorem 1.4 are sharp, and moreover that there are character varieties with arbitrarily large gonality. All of the filled manifolds in the following examples are knot complements in S3S^{3}. For these M(,r)M(-,r), A0(M(,r))A_{0}(M(-,r)) is birational to X0(M(,r))X_{0}(M(-,r)).

9.1. Once-punctured torus bundles of tunnel number one

Refer to caption
Figure 3. The once-punctured torus bundle of tunnel number one, MnM_{n}, is (n+2)-(n+2) surgery on one component of the Whitehead link, shown above.

The family of once-punctured torus bundles of tunnel number one is an infinite family of one-cusped manifolds, indexed as {Mn}\{M_{n}\} for nn\in\mathbb{Z}. These manifolds are hyperbolic for |n|>2|n|>2 and can be realized as integral fillings of one component of the Whitehead link complement, WW; the manifold MnM_{n} is W(,(n+2))W(-,-(n+2)).

The character varieties of these manifolds were explicitly computed in [1]. For each hyperbolic MnM_{n} there is a single canonical component in the SL2()\text{SL}_{2}(\mathbb{C}) character variety. The curves X0(Mn)X_{0}(M_{n}) are all hyperelliptic curves of genus 12(|n1|1)\lfloor\tfrac{1}{2}(|n-1|-1)\rfloor if n2(mod4)n\neq 2\pmod{4} and 12(|n1|3)\lfloor\tfrac{1}{2}(|n-1|-3)\rfloor if n2(mod4)n\equiv 2\pmod{4}. (The variety X0(M3)X_{0}(M_{3}) associated to the figure-eight knot sister is a a rational curve. The variety X0(M3)X_{0}(M_{-3}) associated to the figure-eight knot complement and the variety X0(M6)X_{0}(M_{6}) are elliptic curves.) The discrepancy arising when n2(mod4)n\equiv 2\pmod{4} is due to the fact that in this case there are two components of irreducible representations, the canonical component and an affine line component.

For this family r=p/q=((n+2))/1r=p/q=(-(n+2))/1; the gonality of X0(Mn)X_{0}(M_{n}) is bounded by 2, and the genus is increasing linearly in pp. Note that Theorem 1.2 gives an upper bound of the form cp2c\cdot p^{2} for the genus. The canonical components of the PSL2()\text{PSL}_{2}(\mathbb{C}) character variety are all birational to 1\mathbb{C}^{1} and we conclude that here both the gonality and the genus are bounded by the constant one.

9.2. Double Twist Knots

The double twist knots are the knots J(k,l)J(k,l) shown in the right hand side of Figure 4. These knots are the result of 1/k-1/k surgery and 1/l-1/l surgery on two components of the link shown in the left hand side of Figure 4 where kk and ll are the number of half twists in the boxes. The link J(k,l)J(k,l) is a knot if klkl is even and otherwise is a two component link. Since J(k,l)J(k,l) is ambient isotopic to J(l,k)J(l,k) we may assume that l=2nl=2n. The knot complement of J(k,l)J(k,l) is hyperbolic unless |k||k| or |l||l| is less than two, or k=l=±2k=l=\pm 2. The knots with |k||k| or |l||l| equal to two are the twist knots. These include the figure-eight knot, J(2,2)=J(2,2)=J(2,3)J(2,-2)=J(-2,2)=J(2,3) and the trefoil J(2,2)=J(2,2)J(2,2)=J(-2,-2). We now restrict our attention to these hyperbolic knots.

Let Y0(k,l)Y_{0}(k,l) denote Y0(S3J(k,l))Y_{0}(S^{3}-J(k,l)), and X0(k,l)X_{0}(k,l) denote X0(S3J(k,l))X_{0}(S^{3}-J(k,l)). For a fixed k0k_{0}, by Theorem 1.1

γ(Y0(k0,l))c=c(k0).\gamma(Y_{0}(k_{0},l))\sim c=c(k_{0}).

A similar statement is true if we fix ll.

Refer to caption
Figure 4. The link J(k,l)J(k,l) is the result of 1/k-1/k and 1/l-1/l surgery on the three component link pictured on the left.

The character varieties of these knot complements were computed in and analyzed in [14]. The component of the PSL2()\text{PSL}_{2}(\mathbb{C}) character variety of the double twist knot S3J(k,l)S^{3}-J(k,l) which contain irreducible representations is birationally equivalent to the curve defined by the following equations in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}:

(*) gm+1(r)gn(t)=gm(r)gn+1(t)\displaystyle g_{m^{\prime}+1}(r)g_{n^{\prime}}(t)=g_{m^{\prime}}(r)g_{n^{\prime}+1}(t)  if k=2m,l=2n\displaystyle\ \text{ if }\ k=2m^{\prime},l=2n^{\prime}
fm+1(r)gn(t)=fm(r)gn+1(t)\displaystyle f_{m^{\prime}+1}(r)g_{n^{\prime}}(t)=f_{m^{\prime}}(r)g_{n^{\prime}+1}(t)  if k=2m+1,l=2n.\displaystyle\ \text{ if }\ k=2m^{\prime}+1,l=2n^{\prime}.

(The polynomials fif_{i} and gig_{i} were introduced in Definition 6.1.) These curves are smooth and irreducible in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} when klk\neq l. When k=lk=l the canonical component is given by r=tr=t. The genus can be computed from the bidegree of these polynomials. Let m=|k|2m=\lfloor\tfrac{|k|}{2}\rfloor and n=|l|2n=\lfloor\tfrac{|l|}{2}\rfloor and let D(k,l)D(k,l) be the projective closure of the variety determined by the equations ()(*) in 1(r)×1(t)\mathbb{P}^{1}(r)\times\mathbb{P}^{1}(t). When klk\neq l the bidegree is (m,n)(m,n). The genus of X0(k,l)X_{0}(k,l) is determined by studying the ramification of the map from X0(k,l)X_{0}(k,l) to Y0(k,l)Y_{0}(k,l).

Theorem 1 ([14] Theorems 6.2 and 6.5).

Let J(k,l)J(k,l) be a hyperbolic with m=|k|2m=\lfloor\tfrac{|k|}{2}\rfloor. If klk\neq l the genus of Y0(k,l)Y_{0}(k,l) is (m1)(n1)(m-1)(n-1) and the genus of X0(k,l)X_{0}(k,l) is 3mnmnb3mn-m-n-b where

b={(1)k+l if (1)k+lkl>00 otherwiseb=\left\{\begin{aligned} (-1)^{k+l}&\ \ \text{ if }\quad(-1)^{k+l}kl>0\\ 0&\quad\text{ otherwise}\end{aligned}\right.

The genus of Y0(k,k)Y_{0}(k,k) is zero and the genus of X0(k,k)X_{0}(k,k) is n1n-1.

We will compute the gonality of these curves by relating the gonality to the degree of the curves. In 1883 Noether [18] proved that a non-singular plane curve of degree d2d\geq 2 in 2\mathbb{P}^{2} has gonality equal to d1d-1, where dd is the degree of the curve. (See [16] Theorem 2.3.1, page 73.) Certain well-behaved singularities are known to modify this relationship in an understood manner. The curves determined in [14] are smooth, but lie in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}; the birational mapping to 2\mathbb{P}^{2} introduces singularities in the curves. We now prove a 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} version of Noether’s theorem by understanding this mapping.

Lemma 9.1.

Let CC be a smooth irreducible curve in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} of bidegree (M,N)(M,N) where MN0MN\neq 0, then γ(C)=min{M,N}\gamma(C)=\min\{M,N\}.

(If M=0M=0 or N=0N=0 then CC is isomorphic to 1\mathbb{P}^{1} and therefore has gonality one.)

Proof.

Consider an irreducible curve C2C\subset\mathbb{P}^{2} of degree dd. Let ν\nu be the maximum multiplicity of a singular point of CC. The projection map from the point with multiplicity ν\nu to a line has degree dνd-\nu, so we conclude that γ(C)dν\gamma(C)\leq d-\nu. Ohkouchi and Sakai [19] (Theorem 3) show that if CC has at most two singular points (or infinitely near singular points), then γ(C)=dν\gamma(C)=d-\nu. (See [15] page 194 for a discussion of infinitely near points.)

We consider the birational mapping ϕ:1×12\phi:\mathbb{P}^{1}\times\mathbb{P}^{1}\rightarrow\mathbb{P}^{2} given by

([x:w],[y:z])[wy:zx:xy].([x:w],[y:z])\mapsto[wy:zx:xy].

This has birational inverse given by

ϕ1([a:b:c])=([c:a],[c:b]).\phi^{-1}([a:b:c])=([c:a],[c:b]).

Let DD be a smooth curve of bidegree (M,N)(M,N) in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} with MNM\geq N. It suffices to show that ν=M\nu=M, deg(ϕ(D))=M+Ndeg(\phi(D))=M+N, and ϕ(D)\phi(D) has exactly two singular or infinitely near singular points; for then Ohkouchi and Sakai’s result implies that γ(ϕ(D))=dν=N\gamma(\phi(D))=d-\nu=N. Since ϕ\phi is birational, γ(D)=γ(ϕ(D))\gamma(D)=\gamma(\phi(D)) proving the result.

If DD is a smooth curve of bidegree (M,N)(M,N) then DD is the vanishing set of a polynomial f(x,y)f(x,y) after projectivizing in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. If f(x,y)=cn,mxnymf(x,y)=\sum c_{n,m}x^{n}y^{m} the projectivization of ff is

F(x,w,y,z)=cn,mxnwNnymzMm.F(x,w,y,z)=\sum c_{n,m}x^{n}w^{N-n}y^{m}z^{M-m}.

When x=0x=0 (therefore w=1w=1) this reduces to c0,MyMzMm\sum c_{0,M}y^{M}z^{M-m}. We conclude that there are MM points on DD such that x=0x=0. Similarly there are NN points corresponding to y=0y=0 (and z=1z=1). When the polynomial has a constant term, then these points do not coincide. If ff has no constant term, then ([0:1],[0:1])D([0:1],[0:1])\in D. We will assume, by taking an isomorphic copy of DD, if necessary, that ([0:1],[0:1])D([0:1],[0:1])\not\in D, so ff has a constant term.

Under the birational mapping, ϕ([0:1],[y:z])=[y:0:0]=[1:0:0]\phi([0:1],[y:z])=[y:0:0]=[1:0:0] (if y0y\neq 0). Likewise, ϕ([x:w],[0:1])=[0:x:0]=[0:1:0]\phi([x:w],[0:1])=[0:x:0]=[0:1:0] (if x0x\neq 0). It follows that the NN points above get identified, as do the MM points. The image C=ϕ(D)C=\phi(D) in 2\mathbb{P}^{2} has a singularity of order MM and one of order NN. It is straightforward to see that if xy0xy\neq 0 then a neighborhood of the point ([x:w],[y:z])([x:w],[y:z]) maps smoothly into 2\mathbb{P}^{2}. Therefore ϕ(D)\phi(D) has exactly two singular points, one of degree MM and one of degree NN.

The degree of CC is given by the intersection number CHCH where HH^{\prime} is a hyperplane section of 2\mathbb{P}^{2}. Similarly, the degree of DD is the intersection number DHDH where HH is a hyperplane section of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. The divisor class group of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} is Cl(1×1)=Cl(\mathbb{P}^{1}\times\mathbb{P}^{1})=\mathbb{Z}\oplus\mathbb{Z}. Let E=1×{x}E=\mathbb{P}^{1}\times\{x\} and F={x}×1F=\{x\}\times\mathbb{P}^{1}. Then DD is linearly equivalent to ME+NFME+NF, that is DME+NFD\sim ME+NF. (Two divisors are linear equivalent if their difference is a principal divisor.) The curves of bidegree (M,N)(M,N) correspond to the effective divisors of the class ME+NFME+NF. Similarly, HE+FH\sim E+F.

Therefore, since the degree of a divisor depends only on the linear equivalence class, deg(D)=deg(ME+NF)deg(D)=deg(ME+NF). Since the degree of DD is the intersection number DHDH, and DME+NFD\sim ME+NF and HE+FH\sim E+F it follows that the degree of DD is the intersection number of ME+NFME+NF and E+FE+F, which is M+NM+N.

We now use this to determine the gonality of the curves Y0(k,l)Y_{0}(k,l).

Theorem 9.2.

Let J(k,l)J(k,l) be a hyperbolic knot and let m=12|k|m=\lfloor\tfrac{1}{2}|k|\rfloor and n=12|l|n=\lfloor\tfrac{1}{2}|l|\rfloor.

  • If klk\neq l then γ(Y0(k,l))=min{m,n}\gamma(Y_{0}(k,l))=\min\{m,n\}.

  • If k=lk=l then γ(Y0(k,l))=1\gamma(Y_{0}(k,l))=1.

Proof.

The equations ()(*) defining Y0(k,l)Y_{0}(k,l) are irreducible and determine smooth curves in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} by [14]. These equations have bidegree (m,n)(m,n) if klk\neq l. By Lemma 9.1 if nm0nm\neq 0 the gonality of Y0(k,l)Y_{0}(k,l) is min{m,n}\min\{m,n\}. If m,n=0m,n=0 then k=0,±1k=0,\pm 1 or l=0,±1l=0,\pm 1 and J(k,l)J(k,l) is not hyperbolic. It suffices to consider the case when k=l=2nk=l=2n^{\prime}. In this case, the canonical component for the PSL2()\text{PSL}_{2}(\mathbb{C}) character variety is isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, given by r=tr=t. We conclude that the gonality is one.

Corollary 9.3.

For all positive integers nn there are infinitely many double twist knots KK such that

γ(Y0(S3K))=n.\gamma(Y_{0}(S^{3}-K))=n.

Now we compute the gonality of the canonical components of the SL2()\text{SL}_{2}(\mathbb{C}) character varieties. By Lemma 3.5 since X0(k,l)X_{0}(k,l) is a degree 2 branched cover of Y0(k,l)Y_{0}(k,l) we conclude that

min{m,n}γ(X0(k,l))2min{m,n}.\min\{m,n\}\leq\gamma(X_{0}(k,l))\leq 2\min\{m,n\}.

We now explicitly show that the gonality equals this upper bound.

Theorem 9.4.

Let J(k,l)J(k,l) be a hyperbolic knot and let m=12|k|m=\lfloor\tfrac{1}{2}|k|\rfloor and n=12|l|n=\lfloor\tfrac{1}{2}|l|\rfloor.

  • If klk\neq l then γ(X0(k,l))=2min{m,n}\gamma(X_{0}(k,l))=2\min\{m,n\}.

  • If k=lk=l then γ(X0(k,l))=2\gamma(X_{0}(k,l))=2.

Proof.

We first consider the case when k=l=2nk=l=2n^{\prime}. By [14], a ‘natural model’ for X0(2n,2n)2(x,r)X_{0}(2n^{\prime},2n^{\prime})\subset\mathbb{C}^{2}(x,r) is given by

x2=(r+2)fn(r)21fn(r)2x^{2}=\frac{(r+2)f_{n^{\prime}}(r)^{2}-1}{f_{n^{\prime}}(r)^{2}}

and is birational to the hyper elliptic curve

w2=(r+2)fn(r)21.w^{2}=(r+2)f_{n^{\prime}}(r)^{2}-1.

This has genus n1n^{\prime}-1 by [1]. If J(k,k)J(k,k) is hyperbolic then |n|>2|n^{\prime}|>2 and the gonality of X0(k,k)X_{0}(k,k) is two.

Now, assume that klk\neq l and J(k,l)J(k,l) is hyperbolic. If mm or nn is one, then one can verify directly from the defining equations that X0(k,l)X_{0}(k,l) is an elliptic or hyperelliptic curve and therefore has gonality two. We will now assume that mm and nn are at least two.

Consider the following situation. Let XX and YY be smooth curves of genus gXg_{X} and gYg_{Y} and let f:XYf:X\rightarrow Y be a degree kk simple morphism and h:X1h:X\rightarrow\mathbb{P}^{1} a degree dd morphism. (A simple morphism is one that does not factor through a nontrivial morphism.) The Castelnuovo-Severi inequality (see [3] Theorem 2.1) states that if

(*) dgXkgYk1d\leq\frac{g_{X}-kg_{Y}}{k-1}

then hh factors through ff.

Let XX be a smooth projective model for X0(k,l)X_{0}(k,l) and YY a smooth projective model for Y0(k,l)Y_{0}(k,l). Since there is a degree two birational map from X0(k,l)X_{0}(k,l) to Y0(k,l)Y_{0}(k,l), by Stein factorization, there is a degree two morphism from XX to YY.

From [14], we have a degree k=2k=2 mapping, with gX=3mnmnbg_{X}=3mn-m-n-b (for an explicitly determined b{0,±1}b\in\{0,\pm 1\}) and gY=(m1)(n1)g_{Y}=(m-1)(n-1). Therefore, the righthand side of the inequality is

gX2gY=(3mnmnb)2(m1)(n1)=mn+m+nb2.g_{X}-2g_{Y}=(3mn-m-n-b)-2(m-1)(n-1)=mn+m+n-b-2.

Since b{0,±1}b\in\{0,\pm 1\} the right-hand side of ()(*) is at most mn+m+n1mn+m+n-1. We conclude that any dense map from X0(k,l)X_{0}(k,l) to 1\mathbb{P}^{1} of degree at most mn+m+n1mn+m+n-1 factors through a map from Y0(k,l)Y_{0}(k,l) to 1\mathbb{P}^{1}. Let dd be the minimal degree of a map from X0(k,l)X_{0}(k,l) to 1\mathbb{P}^{1}. Since the gonality of Y0(k,l)Y_{0}(k,l) is min{m,n}\min\{m,n\} we conclude that

min{m,n}d2min{m,n}.\min\{m,n\}\leq d\leq 2\min\{m,n\}.

Since mm and nn are at least two,

d2min{m,n}mn+m+n1.d\leq 2\min\{m,n\}\leq mn+m+n-1.

By the Castelnuovo-Severi inequality, the map factors through Y0(k,l)Y_{0}(k,l). As the gonality of Y0(k,l)Y_{0}(k,l) is min{m,n}\min\{m,n\} by Lemma 9.2 we conclude that the smallest degree of a map from X0(k,l)1X_{0}(k,l)\rightarrow\mathbb{P}^{1} is 2min{m,n}2\min\{m,n\}.

Remark 9.5.

The Brill-Noether theorem gives the bound γg+32\gamma\leq\lfloor\frac{g+3}{2}\rfloor where gg is the genus. For Y0(k,l)Y_{0}(k,l) this inequality is

min{m,n}=γ12(mnmn+2)\min\{m,n\}=\gamma\leq\lfloor\tfrac{1}{2}(mn-m-n+2)\rfloor

so the gonality is smaller than the Brill-Noether bound. For k=lk=l the genus of Y0(k,l)Y_{0}(k,l) is zero and the gonality is one, meeting the Brill-Noether bound. The examples with klk\neq l show that for any NN there is a double twist knot (a twist knot, in fact) such that 12(g+3)γ>N\tfrac{1}{2}(g+3)-\gamma>N. That is, the discrepancy between the gonality and the Brill-Noether bound is unbounded in this family.

Remark 9.6.

Hoste and Shanahan have computed a recursive formula for the AA-polynomials of the twist knots [11], the knots J(±2,n)J(\pm 2,n) and shown that these polynomials are irreducible. Their results show that the degree of the AA-polynomials grow linearly in nn. This shows that the linear bounds in Theorem 1.4 are sharp, since the twist knots are 1/n-1/n surgery on one component of the Whitehead link.

10. Gonality, Injectivity Radius, and Eigenvalues of the Laplacian

Since the smooth projective model of a (possibly singular) affine curve is a compact Riemann surface, it can also be viewed as a quotient of the Poincare disk by a discrete group of isometries of the Poincare disk, and as such it is natural to try to relate gonality to geometric and analytical properties associated to the hyperbolic metric; for example the injectivity radius, ϱ\varrho (i.e. half of the length of a shortest closed geodesic), or the first non-zero eigenvalue of the Laplacian, λ1\lambda_{1}.

To that end, a famous result of Li and Yau [13] shows that for a compact Riemann surface of genus gg, a bound for the gonality γ\gamma is given by:

λ1(g1)γ\lambda_{1}(g-1)\leq\gamma

where γ\gamma is the gonality defined using regular maps. Since the (rational) gonality of a complex affine algebraic curve equals the (regular) gonality of a smooth projective completion, this bound applies to smooth projective models of character varieties. A Riemann surface of genus zero is rational and so has gonality one, and a Riemann surface of genus one is an elliptic curve and so has gonality two. For any higher genus, the above bound implies that

λ1γg1.\lambda_{1}\leq\frac{\gamma}{g-1}.

By Theorem 1.1 the (rational) gonality of X0(M(,r))X_{0}(M(-,r)) is bounded independent of rr, and therefore so is the (regular) gonality, γ\gamma, of a smooth projective model. We conclude that λ1\lambda_{1} is bounded above independent of rr for these curves. Moreover, if {r}\{r\} is sequence of rational numbers such that the genus of the character varieties grows, then since γ\gamma is bounded

λ1γg10.\lambda_{1}\leq\frac{\gamma}{g-1}\rightarrow 0.

For hyperbolic double twist knots the genera of Y0(k,l)Y_{0}(k,l) and X0(k,l)X_{0}(k,l) was determined in [14]. (See § 9.2.) The genus of X0(k,l)X_{0}(k,l) is less than two only for the complements of the knots J(2,2)J(2,-2) (the figure-eight) and J(4,4)J(4,4); both have genus 1. The only hyperbolic double twist knots such that the genus of Y0(k,l)Y_{0}(k,l) is less than two are J(k,k)J(k,k), J(2,2)J(2,-2), J(2,3)J(2,-3) which have genus zero, and J(4,±5)J(4,\pm 5) which have genus one. When klk\neq l, we obtain the following inequalities using Theorem 1

λ1(Y0(k,l))min{m,n}(m1)(n1),λ1(X0(k,l))2min{m,n}3mnmnb\lambda_{1}(Y_{0}(k,l))\leq\frac{\min\{m,n\}}{(m-1)(n-1)},\qquad\lambda_{1}(X_{0}(k,l))\leq\frac{2\min\{m,n\}}{3mn-m-n-b}

where X0(k,l)X_{0}(k,l) and Y0(k,l)Y_{0}(k,l) denote smooth projective models and n=12|k|n=\lfloor\tfrac{1}{2}|k|\rfloor and m=12|l|.m=\lfloor\tfrac{1}{2}|l|\rfloor. Both quantities 0\rightarrow 0 as mm or nn .\rightarrow\infty. For the once-punctured torus bundles of tunnel number one, (see § 9.1) let X0(Mn)X_{0}(M_{n}) denote a smooth projective model. Then for |n|>4|n|>4

λ1(X0(Mn))212(|n2|2)4|n1|3\lambda_{1}(X_{0}(M_{n}))\leq\frac{2}{\lfloor\tfrac{1}{2}(|n-2|-2)\rfloor}\leq\frac{4}{|n-1|-3}

and we conclude that λ10\lambda_{1}\rightarrow 0 as |n||n|\rightarrow\infty.

More recently Hwang and To [12] show that the injectivity radius ϱ\varrho of a compact Riemann surface of genus g2g\geq 2 is bounded above by a function of gonality (defined for regular maps),

ϱ2cosh1(γ).\varrho\leq 2\cosh^{-1}(\gamma).

With X0(M(,r))X_{0}(M(-,r)) denoting a smooth projective model, Theorem 1.1 implies that there is a constant cc such that the gonality of these varieties is bounded above by cc and therefore there is a constant dd such that

ϱ(X0(M(,r)))2cosh1(γ)=d.\varrho(X_{0}(M(-,r)))\leq 2\cosh^{-1}(\gamma)=d.

It follows that the injectivity radius is bounded above for all X0(M(,r))X_{0}(M(-,r)), such that g0,1g\neq 0,1.

By Theorem 9.4 and Theorem 9.2 for all hyperbolic double twist knots other than the figure-eight, if klk\neq l

ϱ(X0(k,l))2cosh1(2min{m,n}),and ϱ(Y0(k,l))2cosh1(min{m,n}).\varrho(X_{0}(k,l))\leq 2\cosh^{-1}(2\min\{m,n\}),\quad\text{and }\quad\varrho(Y_{0}(k,l))\leq 2\cosh^{-1}(\min\{m,n\}).

If k=lk=l then ϱ(X0(k,l))2cosh1(2)\varrho(X_{0}(k,l))\leq 2\cosh^{-1}(2). For all once-punctured torus bundles of tunnel number one with |n|>6|n|>6, X0(Mn)X_{0}(M_{n}) is a hyper elliptic curve. We conclude that for the corresponding smooth projective models the injectivity radius is bounded by 2cosh1(2).2\cosh^{-1}(2).

Motivated by the discussion here, we close with two questions. To state these define an infinite family of compact Riemann surfaces {Σj}\{\Sigma_{j}\} to be an expander family if the genera of Σj\Sigma_{j}\rightarrow\infty and there is a constant C>0C>0 for which λ1(Σj)>C\lambda_{1}(\Sigma_{j})>C. One such family are congruence arithmetic surfaces.
Question: (1) Does there exist an infinite family of 1-cusped hyperbolic 3-manifolds MjM_{j} for which X0(Mj)X_{0}(M_{j}) (resp. Y0(Mj)Y_{0}(M_{j})) is an expander family?

(2) Does there exist an infinite family of 1-cusped hyperbolic 3-manifolds MjM_{j} for which X0(Mj)X_{0}(M_{j}) (resp. Y0(Mj)Y_{0}(M_{j})) has arbitrarily large injectivity radius?

References

  • [1] Kenneth L. Baker and Kathleen L. Petersen, Character varieties of once-punctured torus bundles of tunnel number one, Internat. J. Math. 24 (2013), no. 6, 1350048 (57 pages).
  • [2] S. Boyer and X. Zhang, On Culler-Shalen seminorms and Dehn filling, Ann. of Math. (2) 148 (1998), no. 3, 737–801. MR 1670053 (2000d:57028)
  • [3] Youngook Choi and Seonja Kim, Gonality and Clifford index of projective curves on ruled surfaces, Proc. Amer. Math. Soc. 140 (2012), no. 2, 393–402. MR 2846309 (2012m:14064)
  • [4] D. Cooper and D. D. Long, The AA-polynomial has ones in the corners, Bull. London Math. Soc. 29 (1997), no. 2, 231–238. MR 1426004 (98f:57026)
  • [5] M. Culler and P. B. Shalen, Bounded, separating, incompressible surfaces in knot manifolds, Invent. Math. 75 (1984), no. 3, 537–545. MR 735339 (85k:57010)
  • [6] Marc Culler, Lifting representations to covering groups, Adv. in Math. 59 (1986), no. 1, 64–70. MR 825087 (87g:22009)
  • [7] Nathan M. Dunfield, Cyclic surgery, degrees of maps of character curves, and volume rigidity for hyperbolic manifolds, Invent. Math. 136 (1999), no. 3, 623–657. MR 1695208 (2000d:57022)
  • [8] Jordan S. Ellenberg, Chris Hall, and Emmanuel Kowalski, Expander graphs, gonality, and variation of Galois representations, Duke Math. J. 161 (2012), no. 7, 1233–1275. MR 2922374
  • [9] Robin Hartshorne, Algebraic geometry, Springer-Verlag, New York, 1977, Graduate Texts in Mathematics, No. 52. MR 0463157 (57 #3116)
  • [10] Heisuke Hironaka, Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II, Ann. of Math. (2) 79 (1964), 109–203; ibid. (2) 79 (1964), 205–326. MR 0199184 (33 #7333)
  • [11] Jim Hoste and Patrick D. Shanahan, A formula for the A-polynomial of twist knots, J. Knot Theory Ramifications 13 (2004), no. 2, 193–209. MR 2047468 (2005c:57006)
  • [12] Jun-Muk Hwang and Wing-Keung To, Injectivity radius and gonality of a compact Riemann surface, Amer. J. Math. 134 (2012), no. 1, 259–283. MR 2876146
  • [13] Peter Li and Shing Tung Yau, A new conformal invariant and its applications to the Willmore conjecture and the first eigenvalue of compact surfaces, Invent. Math. 69 (1982), no. 2, 269–291. MR 674407 (84f:53049)
  • [14] Melissa L. Macasieb, Kathleen L. Petersen, and Ronald M. van Luijk, On character varieties of two-bridge knot groups, Proc. Lond. Math. Soc. (3) 103 (2011), no. 3, 473–507. MR 2827003
  • [15] Masayoshi Miyanishi, Algebraic geometry, Translations of Mathematical Monographs, vol. 136, American Mathematical Society, Providence, RI, 1994, Translated from the 1990 Japanese original by the author. MR 1284715 (95g:14001)
  • [16] Makoto Namba, Families of meromorphic functions on compact Riemann surfaces, Lecture Notes in Mathematics, vol. 767, Springer, Berlin, 1979. MR 555241 (82i:32046)
  • [17] Walter D. Neumann and Alan W. Reid, Rigidity of cusps in deformations of hyperbolic 33-orbifolds, Math. Ann. 295 (1993), no. 2, 223–237. MR 1202390 (94c:57025)
  • [18] M. Noether, Zur Grundlegung der Theorie der algebraischen Raumcurven, Abhandlungen der K oniglichen Akademie der Wissenschaften, Berlin (1883).
  • [19] Masahito Ohkouchi and Fumio Sakai, The gonality of singular plane curves, Tokyo J. Math. 27 (2004), no. 1, 137–147. MR 2060080 (2005d:14046)
  • [20] Bjorn Poonen, Gonality of modular curves in characteristic pp, Math. Res. Lett. 14 (2007), no. 4, 691–701. MR 2335995 (2008k:11061)
  • [21] Joan Porti, Torsion de Reidemeister pour les variétés hyperboliques, Mem. Amer. Math. Soc. 128 (1997), no. 612, x+139. MR 1396960 (98g:57034)
  • [22] Igor R. Shafarevich, Basic algebraic geometry. 1, second ed., Springer-Verlag, Berlin, 1994, Varieties in projective space, Translated from the 1988 Russian edition and with notes by Miles Reid. MR 1328833 (95m:14001)
  • [23] Peter B. Shalen, Representations of 3-manifold groups, Handbook of geometric topology, North-Holland, Amsterdam, 2002, pp. 955–1044. MR 1886685 (2003d:57002)
  • [24] W.P. Thurston, The geometry and topology of 3-manifolds, Princeton University, Mathematics Department, 1978, available at http://msri.org/publications/books/gt3m/.
  • [25] Stephan Tillmann, Boundary slopes and the logarithmic limit set, Topology 44 (2005), no. 1, 203–216. MR 2104008 (2005m:57027)
  • [26] Jarosław Włodarczyk, Toroidal varieties and the weak factorization theorem, Invent. Math. 154 (2003), no. 2, 223–331. MR 2013783 (2004m:14113)