JONES-TYPE LINK INVARIANTS AND APPLICATIONS TO 3-MANIFOLD TOPOLOGY By Christine Ruey Shan Lee A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics - Doctor of Philosophy 2015 ABSTRACT JONES-TYPE LINK INVARIANTS AND APPLICATIONS TO 3-MANIFOLD TOPOLOGY By Christine Ruey Shan Lee It is known that the Slope Conjecture is true for an adequate link, and that the colored Jones polynomial of a semi-adequate link has a well-defined tail (head) consisting of stable coefficients, which carry geometric and topological information of the link complement. We study the colored Jones polynomial of a link that is not semi-adequate and show that a tail (head) consisting of stable coefficients of the polynomial can also be defined. Then, we prove the Slope Conjecture for a new family of pretzel knots which are not adequate. We also study the relationship between the Jones polynomial and the topology of the knot complement by relating its coefficients to the non-orientable genus of an alternating knot. The two-sided bound we obtain can often determine the non-orientable genus. Lastly, we develop the connection between the Jones polynomial and a knot invariant coming from Heegaard Floer homology by using it to classify 3-braids which are L-space knots. For my parents iii ACKNOWLEDGMENTS I would like to thank Effie Kalfagianni, for being the advisor that I needed, and the best that I could hope for. She never hesitates to remind me that I can do better. Our conversations are always inspiring, even when my thoughts are dulled. I would like to extend my gratitude to Matt Hedden, for many enlightening discussions and helpful advice. I am also grateful for Ben Schmidt and Bob Bell for reading my writing and offering suggestions. I am happy to acknowledge my friends at/from Michigan State University, in no particular ´ order: Akos, for always knowing more math and being my laughing buddy; Samantha, for being silly together in and out of the office; Charlotte, for being German and smart, and willing to read the introduction of my paper; Allison, who is not in math, but we are all the better for it; Fery, for always knowing when to call; Adam G., for being my academic brother; Adam C. and Andrew, for showing me what life is like after grad school. It would be difficult to imagine surviving grad school without you. Lastly, I am grateful to Ron Fintushel. Without his phone call and well-timed encouragement, I would not be here. iv TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii Chapter 1 Introduction . . . . . . . . . . . . . . . . 1.1 Main results and organization of dissertation . . . 1.1.1 Chapter 2 . . . . . . . . . . . . . . . . . . 1.1.2 Chapter 3 . . . . . . . . . . . . . . . . . . 1.1.3 Chapter 4 . . . . . . . . . . . . . . . . . . 1.1.4 Chapter 5 . . . . . . . . . . . . . . . . . . 1.2 Preliminaries . . . . . . . . . . . . . . . . . . . . 1.2.1 The Temperley-Lieb algebra . . . . . . . . 1.2.2 The Jones-Wenzl idempotent . . . . . . . 1.2.3 Formula for the colored Jones polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 3 3 5 6 8 8 9 10 11 Chapter 2 Stability properties of the colored Jones polynomial 2.1 The colored Jones polynomial of semi-adequate links. . . . . . . . 2.1.1 The Slope Conjecture . . . . . . . . . . . . . . . . . . . . . 2.1.2 The tail of the colored Jones polynomial . . . . . . . . . . 2.2 Ribbon graphs and the Kauffman bracket . . . . . . . . . . . . . . 2.3 A cancellation lemma . . . . . . . . . . . . . . . . . . . . . . . . . n. . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Labeling edges in HA 2.5 Equivalence classes based on G(H). . . . . . . . . . . . . . . . . . 2.6 Combinatorics of G(H) and ribbon graphs . . . . . . . . . . . . . 2.7 Proof of Theorem 2.2.3 . . . . . . . . . . . . . . . . . . . . . . . . 2.8 A worked out example . . . . . . . . . . . . . . . . . . . . . . . . 2.9 Detecting semi-adequacy using the colored Jones polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 15 15 15 19 26 27 39 42 51 52 58 Chapter 3 The Slope Conjecture and 3-string pretzel knots . 3.1 Essential surfaces and branched surfaces . . . . . . . . . . . . . 3.2 Essential surfaces in 2-bridge knot complements . . . . . . . . . 3.3 Curve systems and boundary slopes for Montesinos knots . . . . 3.4 The Hatcher-Oertel algorithm . . . . . . . . . . . . . . . . . . . 3.4.1 Computing the boundary slope from an edgepath system 3.4.2 The 3-strand pretzel knots P (−1/r, 1/s, 1/t). . . . . . . . 3.5 Computation of the Jones Slope . . . . . . . . . . . . . . . . . . 3.6 Proof of Theorem 1.1.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 62 65 70 75 78 79 84 90 v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 4 Crosscap number of alternating links and the 4.1 Augmented links and estimates with normal surfaces . . 4.1.1 An angled polyhedral decomposition. . . . . . . . 4.1.2 Normal surfaces and combinatorial area. . . . . . 4.1.3 Genus estimates for spanning surfaces . . . . . . . 4.2 Crosscap numbers of Alternating links . . . . . . . . . . 4.2.1 State surfaces and a minimum genus algorithm . 4.2.2 Normalization and crosscap number estimates . . 4.3 Proof of Theorem 1.1.3 . . . . . . . . . . . . . . . . . . . 4.4 Calculations of crosscap numbers . . . . . . . . . . . . . 4.4.1 Lower exact bounds . . . . . . . . . . . . . . . . . 4.4.2 Low crossing knots . . . . . . . . . . . . . . . . . 4.5 Generalizations and questions . . . . . . . . . . . . . . . 4.5.1 Non-alternating links . . . . . . . . . . . . . . . . Jones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 95 96 98 102 105 105 113 121 123 123 127 128 128 Chapter 5 The Jones polynomial, 3-braids, and L-space knots . . . . . . 133 5.1 The Jones polynomial, 3-braids, and the Alexander polynomial . . . . . . . . 134 5.2 Proof of Theorem 1.1.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146 vi LIST OF TABLES Table 2.1 Coefficients for the A-adequate knot 10154 [9]. . . . . . . . . . . . . 18 Table 2.2 The tail for a non A-adequate link. . . . . . . . . . . . . . . . . . . 18 Table 2.3 The tables detail the various possibilities for the equivalence classes corresponding to the different cases for a(H) and their corresponding values of G(H), v(H)−k(H), and g(H), when sa is in RΩ1 or Ω0 . The case for when sa is in LΩ1 is analogous. . . . . . . . . . . . . . . . 56 Table 4.1 Knots where the KnotInfo upper bound agrees with our lower bound. The crosscap number is 3. . . . . . . . . . . . . . . . . . . . . . . . 128 vii LIST OF FIGURES Figure 1.1 The 3-string pretzel knot P (−1/10, 1/7, 1/9). . . . . . . . . . . . . . 6 Figure 1.2 Generators of the Temperley-Lieb algebra. . . . . . . . . . . . . . . 9 Figure 1.3 The Jones Wenzl idempotent and its closure in S(R2 ) [78]. . . . . . 10 Figure 1.4 Recursion formula for the Jones-Wenzl Idempotent [78]. . . . . . . . 11 Figure 1.5 Effect of adding a twist [78]. . . . . . . . . . . . . . . . . . . . . . . 11 Figure 1.6 D2 where D is a diagram of the figure 8 knot. . . . . . . . . . . . . 12 Figure 2.1 A- and B-resolutions of a crossing. . . . . . . . . . . . . . . . . . . . 19 Figure 2.2 The figure 8 knot, its all A-resolution and the resulting all-A state graph. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 Figure 2.3 Identical local picture of the A-resolution of the 3-cable of a crossing. 28 Figure 2.4 From left to right: an illustration of the labeling on the edges of en when n = 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 Figure 2.5 Locally, we replace the two segments making up the state circle (in red) and the segment recording the crossing location (dotted) of the A-resolution by that of the segments coming from the B-resolution. 31 From left to right: an all A-state graph with 3 state circles; an edge inclusion representing a spanning sub-graph of the ribbon graph constructed from the all A-state graph; the state graph associated to the spanning sub-graph, which has 2 state circles. . . . . . . . . . . . . 31 On the left: At least one of the two edges smaller than m in Ω1 needs to be included in order that m ends up with two ends on distinct state circles in H . The same is true with the edges bigger than M . On the right, the portion of the region Ω0 cut off by the edge 2n − 1 is shown in yellow, with the edges with one end attached to it. . . . 33 The base case i = 1. 35 Figure 2.6 Figure 2.7 Figure 2.8 . . . . . . . . . . . . . . . . . . . . . . . . . . viii Figure 2.9 Let sa = 2 ∈ RΩ1 for this example. The point of the sequences is to rule out edges in H that may make edges after the edge labeled 2 in RΩ1 have two ends on distinct state circles in H. The sequence for RΩ1 are {bj } = {2, 5, ∅, . . .} and {tj } = {2, ∅, . . .}. . . . . . . . . . 38 On the left: All the edges shown in blue here after the red edge are candidates to be included in G(H), as long as there are no edges above and below after the middle red edge that are included in H. In the middle and on the right: The inclusion of the top and bottom edges in red means that only three edges in blue after red will be included in G(H). This is so that they will still have two ends on the same state circle as shown on the right when we only resolve the crossings corresponding to the red edges. . . . . . . . . . . . . . . . . . . . . . 39 From left to right: An example of edges in G(H) marked in blue; the state graph resulting from including zero edges from G(H); the state graph resulting from choosing one edge from G(H), note that a state circle splits off; the state graph resulting from including two edges from G(H) with two state circles split off. . . . . . . . . . . . . . . 41 Figure 2.12 On the left we have H and on the right we have H . Their local pictures only differ in the inclusion of two edges e1 and e2 . . . . . . 43 Figure 2.13 The edgeset V is shown in black. . . . . . . . . . . . . . . . . . . . 49 Figure 2.14 The leftmost figure depicts a possible set of bk shown in blue, and the edge excluded from bk by the edge in RΩk−1 . Depending on the sequences {tj }, {bj } for RΩk and RΩk+1 , we have at least an extra edge back from bk+1 or bk+2 . . . . . . . . . . . . . . . . . . . . . . 50 Figure 2.15 From left to right: a non A-adequate link diagram, the all-A state graph, and the all A-state graph of the 3-cable. . . . . . . . . . . . . 53 Figure 3.1 A compressing disk [52]. . . . . . . . . . . . . . . . . . . . . . . . . . 63 Figure 3.2 A ∂-compressing disk [52]. . . . . . . . . . . . . . . . . . . . . . . . 63 Figure 3.3 Models of a branched surface and a branched surface neighborhood [54]. 64 Figure 3.4 Schematics for a Montesinos knot. . . . . . . . . . . . . . . . . . . . 65 Figure 3.5 Figure from [56]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 Figure 3.6 Intersection of S with Sr2 [56]. . . . . . . . . . . . . . . . . . . . . . 68 Figure 2.10 Figure 2.11 ix Figure 3.7 The 1-simplex on the Poincare disc model [56]. . . . . . . . . . . . . 69 Figure 3.8 The generators a, b and c and the corresponding set of disjoint curves on the 4-punctured sphere with a, b, c-coordinates (3, 1, 2). . . . . . 71 Figure 3.9 The three-simplex of curve systems on a four-punctured sphere [55]. 72 Figure 3.10 The four-simplex of curve systems with systems of slope 1/0 added [55]. 73 Figure 3.11 On the left: the infinite strip containing the 1-dimensional complex D. On the right: the portion of the infinite strip from height 0/1 to 1/1 [55]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 Figure 3.12 A choice of edgepath for each of the three fractions 1/2, 1/8, and 3/4 are shown in red, blue, and green. . . . . . . . . . . . . . . . . . . . 80 Figure 3.13 A trivalent graph representation [78]. . . . . . . . . . . . . . . . . . 84 Figure 3.14 The fusion formula [78]. . . . . . . . . . . . . . . . . . . . . . . . . . 85 Figure 3.15 The untwisting formula [78]. . . . . . . . . . . . . . . . . . . . . . . 85 Figure 3.16 The 6j-equation [78]. . . . . . . . . . . . . . . . . . . . . . . . . . . 86 Figure 3.17 Replacing a region bounded by two edges. δad is the Kronecker delta function [78]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 Figure 3.18 Collapsing a triangular region to a vertex [82]. . . . . . . . . . . . . 87 Figure 3.19 The polytope cut out by equations α + β = γ, α + γ = β, β + γ = α. It has been rescaled by dividing by n. . . . . . . . . . . . . . . . . . 89 Figure 4.1 Twist reduced: A or B must be a string of bigons [37]. . . . . . . . . 95 Figure 4.2 From left to right: A link K, an augmented link J and a fully augmented link L [32]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 Figure 4.3 Two ways of augmenting a twist region with a single crossing. . . . . 97 Figure 4.4 Normal disks in a truncated polyhedron [37]. . . . . . . . . . . . . . 99 Figure 4.5 The two resolutions of a crossing, the arcs recording them and their contribution to state surfaces. . . . . . . . . . . . . . . . . . . . . . 106 x Figure 4.6 One branch of the algorithm resolves the crossings so that the triangle becomes a state circle. The other branch resolves them the opposite way. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Figure 4.7 A diagram of the figure 8 knot with bigon regions labeled 1 and 2 and the diagram resulting from applying the first step of Algorithm 4.2.2. The new choices of bigon regions are labeled 1, 2, and 3. . . . 109 Figure 4.8 The state surfaces resulting from resolving the rest of the bigon regions with different choices of bigon regions. . . . . . . . . . . . . . 110 Figure 4.9 The portion of S through a twist region with more than one crossing and the crossing circle for the twist region. . . . . . . . . . . . . . . 111 Figure 4.10 Surfaces from two branches of the splitting in Step 2b in the AdamKindred algorithm and the corresponding choice of augmentation. . 112 Figure 4.11 The knots 103 (left) and 10123 (right) [19]. . . . . . . . . . . . . . . 119 Figure 4.12 From left to right: The portion of D(K) around a vertex of G, the corresponding potion of the augmented link and portion of the polyhedral decomposition. An ideal triangle is indicated by the red line. Figure 5.1 125 The generator σi . Note that this convention is the opposite of that of [14]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 xi Chapter 1 Introduction The colored Jones polynomial belongs to a family of quantum invariants constructed from the ideas of representation theory and physics following the discovery of the Jones polynomial [67,68]. Despite its efficacy at distinguishing knots, its relationship to classical knot topology and the geometry of hyperbolic knots remains to be determined. Several conjectures are open, such as the Slope Conjecture [41], which relates the polynomial to incompressible surfaces in the knot complement, and the celebrated Volume Conjecture [87, 88], which connects the values of the polynomial to the volume of the hyperbolic knot. As much of this dissertation is motivated by the Slope Conjecture, we give its statement here. The colored Jones polynomial assigns to a knot K in S 3 a sequence of Laurent polynon (q)}∞ , where J n (q) ∈ Z[q −1/2 , q 1/2 ]. Let d(n) and d∗ (n) denote the minimum mials {JK n=2 K n (q). The formula for J n (q) based on the Kauffman degree and maximum degree in q of JK K bracket gives a lower bound hn (D) ≤ d(n), and an upper bound Hn (D) ≥ d∗ (n), which can be explicitly computed from any diagram D of K, see Definition 2.2.2. A fraction p/q ∈ Q ∪ {1/0} is a boundary slope of a knot K ⊂ S 3 if pµ + qλ represents the homology class of the boundary of a properly embedded, essential surface in the torus boundary of the knot complement. Here the set {µ, λ} is the canonical meridian-longitude basis of homology of the torus boundary. Garoufalidis’ Slope Conjecture is the following statement. 1 Conjecture 1.0.1. [41] For a knot K ⊂ S 3 , every limit point of each of the sets jsK := 4d(n) :n≥0 n2 js∗K := 4d∗ (n) :n≥0 n2 is a boundary slope of K. The limit points of the sets jsK and js∗K are called the Jones slopes of K, and the conjecture predicts that every Jones slopes is a boundary slope. By the work of Garoufalidis [41] and Garoufalidis and Le [43], the set of Jones slopes for a knot is a finite set of rational numbers. Hatcher [54] showed that there are only finitely many boundary slopes for a link. Therefore, the set of boundary slopes is a way to index geometric information about a knot. Deep relationships between the colored Jones polynomial and the geometry of the knot complement in the spirit of these conjectures have been demonstrated and studied extensively on the class of semi-adequate links [9–11, 25, 33–35, 44, 45, 47, 105]. These are a class of links satisfying a diagrammatic condition. In contrast, very little is known outside the class of semi-adequate links. In the dissertation, we investigate the structure of the colored Jones polynomial outside the class of semi-adequate links. Our first theorem generalizes a stability result known for semi-adequate links to all links by studying the combinatorics of the link diagram. Our second result, joint with R. van der Veen, proves the Slope Conjecture for a class of 3-string pretzel knots. In a different direction, we study the connection between the Jones polynomial and other link invariants. With E. Kalfagianni, we give two-sided linear bounds of the crosscap number (i.e. the non-orientable genus) of an alternating link in terms of coefficients of the Jones polynomial of the link. 2 A rational homology 3-sphere Y is an L-space if |H1 (Y ; Z)| = rank HF (Y ), where HF (Y ) denotes the ‘hat’ version of Heegaard Floer homology. A knot in S 3 is an L-space knot if K or its mirror image admits a positive L-space surgery. L-space knots have particularly simple Heegaard Floer homology groups and are extensively studied for their relationship to len-space knots. Using the Jones polynomial in collaboration with F. Vafaee, we classify the L-space knots among 3-braids. For the rest of the introduction, we give precise statements of the main results of this dissertation in Section 1.1 and describe the contents of each chapter. In Section 1.2, we give preliminaries on the colored Jones polynomial that are essential for the rest of the dissertation. 1.1 Main results and organization of dissertation Each chapter of the dissertation is self-contained and may be read out of order by a reader familiar with the preliminary background on the Jones polynomial and the colored Jones polynomial, which we cover in Section 1.2. 1.1.1 Chapter 2 We say that a link is A-adequate (resp. B-adequate) if it admits an A-adequate (resp. Badequate) diagram, see Definition 2.2.1. A link is semi-adequate if it is A-or B-adequate. It is adequate if it is both A-and B-adequate. As we will see in Chapter 2, the colored Jones polynomial of semi-adequate knots is closely related to the hyperbolic structure of the knot complement. The Slope Conjecture is true for adequate knots [33], and the coefficients of the polynomial give volume bounds for 3 a hyperbolic knot [25, 34], as well as exhibit a stability propery [9, 44]. A key feature of a semi-adequate link is the diagrammatic condition in its definition, which greatly reduces the complexity of computing the polynomial. Since the Slope Conjecture and the Volume Conjecture are expected to hold for all knots, n (q)} satisfies a recurrence relation [43], analogous and it is known that the sequence {JK results are expected to hold for non semi-adequate knots. In particular, we would like to understand how the relationship to the geometry of the knot complement may be different for knots that are not semi-adequate. A way to approach this problem is to study how the lack of the diagrammatic feature of semi-adequacy affects the polynomial, since diagrammatic properties of links are more directly related to the geometry of the link complement. n (q) and h (D) is a lower bound of d(n) Recall that d(n) is the minimal degree of JK n obtained from a diagram D of the link K, see Definition 2.2.2. The main result of Chapter 2 gives a new lower bound for d(n) for a link that is not A-adequate. Theorem 1.1.1. Let K be a link with diagram D. If D is not A-adequate, then d(n) ≥ hn (D) + n − 2 for n > 2. Theorem 1.1.1 is a first step in understanding the structure of the colored Jones polynomial outside the class of semi-adequate links. A immediate consequence of the theorem is that we may define stable coefficients for non A-adequate links which are analogous to the stable coefficients of the polynomial of A-adequate links. In addition, these coefficients are trivial for non A-adequate links. Therefore, the colored Jones polynomial can characterize A-adequate links. To prove Theorem 1.1.1, we use the ribbon graph construction of the Kauffman bracket 4 and a formula of the colored Jones polynomial in terms of the blackboard cable of the link diagram. The computation for the lower bound of the degree is localized to a portion of the diagram that is not A-adequate. 1.1.2 Chapter 3 Outside the class of adequate knots, the Slope Conjecture is often verified by computing the Jones Slopes and boundary slopes separately, and then matching them together [41, 46]. The question of how the colored Jones polynomial selects the boundary slope of an essential surface remains a major obstacle in understanding the conjecture. In particular, there is a lack of examples which would provide an intrinsic explanation for the connection between n (q)} and incompressible surfaces. the degrees of {JK In this chapter, we prove the Slope Conjecture for a new family of pretzel knots. This provides a large number of new examples to study. A 3-string pretzel knot P (−1/r, 1/s, 1/t) has a diagram obtained by connecting a negative twist region with r twists and two positive twist regions with s and t twists, respectively, see Figure 1.1 Theorem 1.1.2 (Lee-van der Veen). Let r, s, t be nonzero integers greater than 5 such that r is even, s, t are odd, and r > s, t. The Slope Conjecture is true for pretzel knots P (−1/r, 1/s, 1/t). As an example, for P (−1/10, 1/7, 1/9), the boundary slope picked out by the colored Jones polynomial in the sense of the Slope Conjecture has slope 302/7 and Euler characteristic -56. The diagram given for P (−1/r, 1/s, 1/t) is A-adequate but not B-adequate, therefore, n (q). Most of the we need only to verify the conjecture for the maximum degree d∗ (n) of JK 5 Figure 1.1: The 3-string pretzel knot P (−1/10, 1/7, 1/9). boundary slopes for each r, s, t matched by the Jones slope of Theorem 1.1.2 are rational. Therefore, we also have many new examples of pretzel knots which are not B-adequate, since the slopes would be integral otherwise [33]. For Theorem 1.1.2, the Hatcher-Oertel algorithm [55] determines the set of boundary slopes for any Montesinos knot. We use this algorithm to match a boundary slope with the Jones slope for the class of pretzel knots in the theorem. 1.1.3 Chapter 4 Let JK (t) = αK tr + βK tr−1 + . . . + βK ts+1 + αK ts denote the Jones polynomial of K, so that r and s denote the highest and lowest power in t. Set TK := |βK | + βK , where βK and βK are the second and penultimate coefficients of JK (t), respectively. 6 The crosscap number of a non-orientable surface with k boundary components is defined to be 2 − χ(S) − k. The crosscap number of a link K is the minimum crosscap number over all non-orientable surfaces spanning K. Theorem 1.1.3 (Kalfagianni-Lee). Let K be a non-split, prime alternating link with kcomponents and with crosscap number C(K). Suppose that K is not a (2, p) torus link. We have TK 3 + 2 − k ≤ C(K) ≤ TK + 2 − k, where TK is as above, and · is the ceiling function that rounds up to the nearest integer. Furthermore, both bounds are sharp. Theorem 1.1.3 gives a new relation of the Jones polynomial to a fundamental topological knot invariant and prompts several interesting questions about the topological content of quantum link invariants. The lower bound is easy to compute from any alternating diagram of the knot. Combined with the information on KnotInfo [19], the lower bound determines previously unknown values of the crosscap numbers of many alternating links. In the proof of Theorem 1.1.3 we make use of the results of [1], which give an algorithm to determine crosscap number of any alternating link. We show that given an alternating diagram D(K), a surface that realizes the crosscap number of K can be taken to lie in the complement of an augmented link obtained from D(K). Then, we use results of Agol, D. Thurston [76, Appendix], Lackenby [75] and Futer and Purcell [37], to prove Theorem 1.1.3. In particular, we make use of the fact that augmented link complements admit angled polyhedral decompositions with several nice combinatorial and geometric features. We employ normal surface theory and a combinatorial version of a Gauss-Bonnet theorem to estimate 7 the Euler characteristic of surfaces that realize crosscap numbers of alternating links. Using these techniques we show that for prime alternating links, the crosscap number is bounded in terms of the twist number of any prime, twist-reduced, alternating projection, see Section 4.1 for definitions. 1.1.4 Chapter 5 One of the most prominent problems in relating low-dimensional topology and Heegaard Floer homology is to give a topological characterization to L-spaces and L-space knots. As an application of the Jones polynomial, we show that we can use it to determine which 3-braids are L-space knots. Theorem 1.1.4 (Lee-Vafaee). Twisted (3, q) torus knots are the only knots with 3-braid representations that admit L-space surgeries. Our proof of Theorem 1.1.4 uses the constraints on the Alexander polynomial of an Lspace knot that have previously been studied to give the classification of L-space knots among pretzel knots [80]. We show that, except for the twisted (3, q) torus knots, the Alexander polynomials of all of the knots with 3-braid representations violate the constraints mentioned for the Alexander polynomial of L-space knots. 1.2 Preliminaries In this section we give the formula of the colored Jones polynomial, and we show how the formula in terms of blackboard cables of diagrams can be derived from it. The colored Jones polynomial may be defined via the representation theory of sl2 (C), see [99], [100], [109], and the references therein. We use the definition of the colored Jones polynomial in terms of 8 the Kauffman bracket and the Jones-Wenzl idempotent of the Temperley-Lieb algebra, a description of which may be found in [78]. 1.2.1 The Temperley-Lieb algebra Let A be a fixed complex number. The linear skein S(F ) of a surface F is the vector space of formal linear sums over C of link diagrams in F quotiented by the relations (i) D ∪ (ii) Here = (−A2 − A−2 )D. =A + A−1 . is regarded as a trivial closed curve in the diagram with no crossing. Consider the linear skein S(D, 2n) of a disc regarded as a square with 2n points marked on its boundary and distinguish one side with n marked points as the ‘left side’ and another side, also with n marked points, as the ‘right side.’ We define a product d1 d2 of two elements d1 , d2 in S(D, 2n) by juxtaposing the squares of d1 and d2 side by side and identifying the right side of d1 with the left side of d2 . Extending this multiplication by linearity, we obtain a bilinear map S(D, 2n) × S(D, 2n) → S(D, 2n) which makes S(D, 2n) into an algebra T Ln , called the Temperley-Lieb algebra, over C. T Ln is generated by n elements 1, e1 , . . . , en−1 . Here the convention is that a strand decorated by a number n indicates n parallel strands. 1= n ei = n−i+1 i−1 Figure 1.2: Generators of the Temperley-Lieb algebra. There is a map from T Ln into S(R2 ) by placing an element of T Ln in the plane and joining the n points on the left side and n points on the right side by parallel arcs. In the 9 linear skein of R2 , the image is reduced to an empty diagram with coefficient a polynomial in A. 1.2.2 The Jones-Wenzl idempotent The Jones-Wenzl idempotent is an element fn in T Ln defined by the following four properties. We represent it by an empty box with n incoming strands and n outgoing strands. n fn = n n = Figure 1.3: The Jones Wenzl idempotent and its closure in S(R2 ) [78]. Let n denote the polynomial multiplying the empty diagram of the image of fn in S(R2 ) obtained by joining the n points on both sides by n parallel strands. Definition 1.2.1. The Jones-Wenzl idempotent fn is the unique element in T Ln which satisfies (i) fn ei = 0 = ei fn for 1 ≤ i ≤ n − 1. (ii) fn − 1 belongs to the algebra generated by {e1 , . . . , en−1 }. (iii) fn fn = fn , and (iv) n = (−1)n (A2(n+1) −A−2(n+1) ) . (A2 −A−2 ) The existence of this element is proven in [78]. An important property of fn is a recursion formula, which we show diagrammatically. 10 1 1 − n−1 = n+1 n n n n−1 n Figure 1.4: Recursion formula for the Jones-Wenzl Idempotent [78]. Adding a kink via a Type I Reidemeister move to the n-strands of f n multiplies the skein element by a scalar [78, Lemma 14.2]. 2 = (−1)n An +2n Figure 1.5: Effect of adding a twist [78]. 1.2.3 Formula for the colored Jones polynomial If D is an unoriented diagram of a link K with k ordered components, D defines a multilinear map , . . . , D. , . . . , D : S(S 1 × I) × · · · × S(S 1 × I) → S(R2 ), k times from the Cartesian product of linear skeins of the annulus S 1 × I to S(R2 ) in the following way: For each component of the link diagram we obtain an annulus corresponding to the blackboard framing, where the boundaries of the annulus run parallel to the diagram as shown for the figure 8 knot in Figure 1.6. We identify S 1 × I with the annulus and thus obtain a map from a skein element S 1 × I into the annulus corresponding to a link component of D. The n + 1th-unreduced colored 11 Figure 1.6: D2 where D is a diagram of the figure 8 knot. n+1 (q) is the polynomial obtained by substituting q 1/2 = A−1/2 into the Jones polynomial JK polynomial multiplying the empty diagram of 2 (−1)n An +2n −ω(D) n, . . . , n ∈ S(R2 ). (1.1) By induction and the recursion formula of fn depicted by Figure 1.4, We have that ,..., n, . . . , = ,...,α n − n−1 , . . . , , where α is the generator of S(S 1 × I) consisting of a parallel curve encircling the annulus. n+1 (q) is also obtained by substituting q 1/2 = A−1/2 Extending by bilinearity, we have that JK into the polynomial GD (n + 1, A). GD (n + 1, A) := 2 (−1)n An +2n −ω(D) Sn (D) , (1.2) where Sn (D) is defined recursively by Sn+1 (D) = DSn (D) − Sn−1 (D). We set S0 (D) = 1 to be 1 times the empty diagram, while S1 (D) = D. Note that when 12 1+1 (q) = J 2 (q) is the Jones polynomial J (q) multiplied by (−A2 − A−2 ). n = 1, we have JK K K Remark 1.2.2. The polynomials {Sn (x)} defined by the recurrence relation are Chebyshev’s polynomials of type 2 [77]. 13 Chapter 2 Stability properties of the colored Jones polynomial The colored Jones polynomial of a link can be defined and studied in terms of combinatorial properties of link diagrams. One approach to this is through the Kauffman bracket construction [72] as we have seen in Section 1.2. Studying the combinatorics of the Kauffman bracket, Lickorish and Thistlethwaite [79] showed that the extreme degrees of the Jones polynomial are bounded by concrete data from any diagram. If the diagram is adequate, which means that it is A-adequate and B-adequate, see Definition 2.2.1, then the bounds are sharp. This was one of the very first results utilizing the notion of semi-adequacy. We summarize below the results known for semi-adequate links that are relevant to this chapter. A diagram is B-adequate if its mirror image is A-adequate, and the effect of taking the mirror image of the diagram on the colored Jones polynomial is to substitute q −1 for q. Therefore, for the rest of the chapter, we will only refer to A-adequacy. Recall that d(n) is n (q). the minimum degree of JK 14 2.1 2.1.1 The colored Jones polynomial of semi-adequate links. The Slope Conjecture The degrees of the colored Jones polynomial, and hence the Jones slopes, are difficult to compute in general. However, if a knot is adequate, the computation boils down to determining hn (D), see Definition 2.2.2. If D is A-adequate, then d(n) = hn (D) [33]. Futer, Kalfagianni, and Purcell [33] use this fact and the results on essential surfaces in the complement of semi-adequate knots to prove the Slope Conjecture for adequate knots. 2.1.2 The tail of the colored Jones polynomial n (q) exhibit a stability behavior for semi-adequate It is also known that the coefficients of JK links. The stable coefficients directly relate to the geometry of hyperbolic knot complements. i (q). If K Theorem 2.1.1. [9, 44] For i ≥ 2, let βi be the coefficient of q d(i)+(i−1) of JK n (q) is admits an A-adequate diagram, then for all n ≥ i, the coefficient of q d(n)+(i−1) of JK equal to βi . n (q) from h (D) are stable for a link with an That is, the last n − 1 coefficients of JK n A-adequate diagram D, extending the results by Dasbach and Lin [25] on the last and n (q). Based on this result, they defined the notion of a penultimate stable coefficients of JK tail of the colored Jones polynomial of an A-adequate link. Definition 2.1.2. For a link admitting an A-adequate diagram, the power series ∞ TK (q) = i=2 15 βi q i is a tail of the colored Jones polynomial. n (q), and hence the last two coefficients of the The last and penultimate coefficients of JK tail, have been shown by the work of Futer, Kalfagianni, and Purcell to carry information on the geometric structure of the complement of an A-adequate knot, and to provide sharp volume bounds on the complement of hyperbolic links. See [34] for the results and a detailed survey. Rozansky has shown that stability behavior also occurs in the categorification of the colored Jones polynomial [101]. The idea of this theory, developed independently by Frenkel, Stroppel and Sussan, Cooper and Krushkal, and Rozansky, is as follows: Given a link K, for each n, one assigns a chain complex C Kh (K, n) for which the graded Euler characteristic of n (q). its homology groups H Kh (K, n) is the n-th colored Jones polynomial JK Theorem 2.1.3. [101] If K is an A-adequate link, then there is a sequence of maps ˜ Kh (K, n) → H ˜ Kh (K, n + 1), fn : H ˜ Kh for i ≤ n − 1, where i is the grading of the chain such that fn is an isomorphism on H i complex coming from the number of state circles of a Kauffman state. The tilde indicates the appropriate degree shifts made for the chain complex so that fn is a degree-preserving map for each n. In other words, for all n > i, the homology groups of C Kh (K, n) of grading less than i − 1 are the same as the homology groups of C Kh (K, i) of grading less than i − 1. As a result, he defines a tail homology of the colored Jones ˜ Kh (K, n) determined by fn . polynomial which is the direct limit of the direct system of H The tail homology contains the stable homology groups of C Kh (K, n) as n increases. The 16 last n − 1 coefficients of the graded Euler characteristic of the tail homology agree with the n (q)). Therefore, last n − 1 coefficients of TK (q) (hence the last n − 1 stable coefficients of JK we have a categorification of the tail in the case of A-adequate links, extending the results by Armond [9] and Garoufalidis and Le [44]. It is natural to ask whether similar results can be obtained outside the class of semiadequate links. Specifically, we are interested in the following questions. Question 2.1. In view of Conjecture 1.0.1, what can we say about the degrees of the colored Jones polynomial when the link is not semi-adequate? Question 2.2. Can we define a tail for the colored Jones polynomial for all links as in Definition 2.1.2? Question 2.3. What is the tail homology as defined in Theorem 2.1.3 for the categorification of the colored Jones polynomial for non semi-adequate links? For Question 2.1, Manchon [81] has constructed an infinite family of non A-adequate knots with diagrams D for which d(2) = h2 (D). An example of such a knot is 12n706, see Knotinfo [19] for a diagram of the knot. For this knot, d(2) = h2 (D) = −4. However, this knot is not A-adequate since the last coefficient of its Jones polynomial is 2, and it is known that for an A-adequate link, the last coefficient of its Jones polynomial is ±1 [25]. In n (q)}∞ is q-holonomic [43]. In general, Garoufalidis and Le showed that the sequence {JK n=2 other words, the entire sequence is determined by finitely many initial terms. This can be viewed as a weaker form of stability in view of Question 2.2. We study the effect that a diagram being non A-adequate has on the last n−2 coefficients n (q). We reprint Theorem 1.1.1 for the convenience of the reader. of JK 17 Theorem 1.1.1. Let K be a link with diagram D. If D is not A-adequate, then d(n) ≥ hn (D) + n − 2 for n > 2. A special case of Theorem 1.1.1 is proven in [69]. If K is not A-adequate, then any link diagram D = D(K) will not be A-adequate. Let h(n) be the maximum of hn (D) taken over all diagrams D of K. By Theorem 1.1.1, d(n) ≥ h(n) + n − 2 for n > 2. We use this to A (q), which is constant if K is not obtain a link invariant in the form of a power series JK A-adequate as in [69]. In other words, the stable coefficients βi are identically zero for i > 2. This allows us to extend the construction of a tail of an A-adequate link to all links. Figure 2.1 and 2.2 illustrate the difference in coefficients of the tail of an A-adequate n (q) from links and non A-adequate links. The symbol a(d(n) + i) is the ith-coefficient of JK d(n) and a(h(n) + i) is similarly defined. If K is A-adequate, then d(n) = h(n). n=2 n=3 n=4 n=5 n=6 a(d(n)) a(d(n) + 1) 1 -2 1 -2 1 -2 1 -2 1 -2 a(d(n) + 2) 2 -1 -1 -1 -1 a(d(n) + 3) -3 5 2 2 2 Table 2.1: Coefficients for the A-adequate knot 10154 [9]. n=2 n=3 n=4 n=5 n=6 a(h(n)) a(h(n)+1) a(h(n)+2) a(h(n)+3) 0 0 0 0 0 0 0 0 0 0 Table 2.2: The tail for a non A-adequate link. Thereby, we give an affirmative answer to Question 2.2 and make partial progress on understanding Question 2.1 and 2.3. With respect to the tail homology, Rozansky made the following conjecture: 18 Conjecture 2.1.4. If K is not A-adequate, then the tail homology for the categorification of the colored Jones polynomial for K vanishes. Thus, the last n − 2-coefficients of the nth-colored Jones polynomial of a non A-adequate link should be identically 0, whence Theorem 1.1.1 provides partial evidence towards Conjecture 2.1.4. We cover the preliminaries on ribbon graphs in Section 2.2. We reduce the theorem to proving Theorem 2.2.3, phrased entirely in terms of the Kauffman bracket of a link diagram. We organize the intermediary lemmas between Section 2.3 - Section 2.6, prove Theorem 2.2.3 in Section 2.7 and show how the algorithm in the proof works on an example in Section 2.8. In Section 2.9, we discuss applications of Theorem 1.1.1, including the construction of a tail for all links, and how the colored Jones polynomial can be used to detect semi-adequate links. 2.2 Ribbon graphs and the Kauffman bracket We review the notion of A-adequacy, which we will use to obtain an upper bound of d∗ Sn−1 (D) . This will be the lower bound hn (D) of d(n) in the statement of Theorem 1.1.1 after substituting q 1/2 = A−1/2 . Let D be an oriented link diagram and x a crossing of D. Associated to D and x are two link diagrams, each with one fewer crossing than D, called the A-resolution and B-resolution of the crossing, see Figure 2.1. A-resolution B-resolution Figure 2.1: A- and B-resolutions of a crossing. 19 A Kauffman state σ of D is a choice of A-resolution or B-resolution at each crossing of D. Applying a Kauffman state to D, we obtain a crossing-free diagram consisting of a disjoint collection of simple closed curves on the projection plane, see Figure 2.2. We call these curves state circles. The all-A state chooses the A-resolution at each crossing of D. We denote the union of the corresponding state circles by sA (D), and let |sA (D)| be the number of state circles. For an arbitrary state σ, we similarly denote the union of the corresponding state circles by sσ (D) and let |sσ (D)| be the number of state circles in the union. Figure 2.2: The figure 8 knot, its all A-resolution and the resulting all-A state graph. Definition 2.2.1. To the union of state circles sσ (D) resulting from applying a Kauffman state σ, we attach one edge for each crossing which records the original location of the crossing on D. (These edges are dashed in Figure 2.1.) We will call the union of the state circles and the edges the graph of the σ-resolution, or the σ-state graph, and denote it by Hσ (D). If σ is the all-A state then we call it the graph of the A-resolution, or the all-A state graph and denote it by HA (D). A diagram D is A-adequate if HA (D) does not contain any edges with both ends on the same state circle. Whenever it is clear which diagram we are referring to, we will just write Hσ for Hσ (D) and HA for HA (D) to simplify notation. Definition 2.2.2. Let c(D) be the number of crossings of D, ω(D) be the writhe of D, and 20 |sA (D)| be the number of state circles in the all A-state of D. We define M (D) := c(D) + 2|sA (D)|, and 1 hn (D) := − (M (Dn−1 ) + 4 − 2(n − 1) − ω(D)((n − 1)2 + 2(n − 1)). 4 It is known that for any link diagram D, the maximum degree d∗ D of D in A is less than or equal to M (D). Moreover, d∗ D = M (D) if D is A-adequate [79]. n+1 (q) is obtained by substituting q 1/2 = A−1/2 into Recall that JK 2 (−1)n An +2n −ω(D) Sn (D) . (2.1) An important corollary of the recursive formula for Sn (D) is that Sn (D) = Dn + (1 − n) Dn−2 + lower degree cablings of D . Therefore, d∗ Sn (D) ≤ M (Dn ) and d(n) ≥ hn (D) for all n > 2. This shows that hn (D) is a lower bound for d(n). If D is A-adequate, then Dn is also A-adequate and thus hn (D) = d(n) [33]. We shall prove an upper bound on the maximum degree of Dn , which will imply Theorem 1.1.1. Theorem 2.2.3. Let D be a link diagram that is not A-adequate. Then for n > 2, we have d∗ Dn ≤ M (Dn ) − 4(n − 1). 21 We show that Theorem 2.2.3 implies Theorem 1.1.1: Proof. The n-cable Dn has n2 c(D) crossings and n|sA (D)| number of state circles in the all-A state, therefore M (Dn ) = n2 c(D) + 2n|sA (D)|. Note that M (Dn ) − M (Dn−2 ) = c(D)(n2 − (n − 2)2 ) + |sA (D)|(2n − 2(n − 2)) > 4(n − 1). Assuming Theorem 2.2.3, so M (Dn ) − d∗ Dn ≥ 4(n − 1), then M (Dn ) − d∗ Sn (D) ≥ 4(n − 1). Plugging this back into the expressions for d(n) and hn (D) gives us that d(n) ≥ hn (D) + n − 2. The Kauffman bracket of any diagram can be computed by means of ribbon graph expansions. As defined in [24], a ribbon graph is a multi-graph (i.e. loops and multiple edges are allowed) equipped with a cyclic order on the edges at every vertex. A ribbon graph can be embedded on an orientable surface such that every region in the complement of the graph is a disk. We called the regions the faces of the ribbon graph. 22 Definition 2.2.4. For a ribbon graph G we define the following combinatorial quantities: v(G) = the number of vertices of G. e(G) = the number of edges of G. f (G) = the number of faces of G. k(G) = the number of connected components of G. g(G) = 2k(G) − v(G) + e(G) − f (G) , the genus of G. 2 For each Kauffman state σ of a link diagram, a ribbon graph Gσ is constructed as follows: We orient the collection of non-intersecting state circles sσ (D) according to whether a circle is inside an odd or even number of circles, respectively. The vertices of Gσ correspond to the collection of state circles and the edges to the crossings. The orientation of the circles defines the orientation of the edges around vertices. An equivalent definition of an Aadequate diagram is to say that a diagram is A-adequate if the ribbon graph GA constructed from the all-A state of the diagram has no one-edge loops. For more details, see [23], where the word dessin is used instead of ribbon graph. Definition 2.2.5. A spanning sub-graph H ⊆ GA is a ribbon graph obtained by removing edges from GA . From the ribbon graph setting we obtain the spanning sub-graph expansion of the Kauffman bracket, originally introduced in [23] and proven in [24]. Theorem 2.2.6. Let GA be the ribbon graph constructed from the all-A state of a link diagram D. Then Ae(GA )−2e(H) (−A2 − A−2 )f (H) , D = H⊆GA 23 where H ranges over all spanning sub-graphs of GA . The term XH defined by XH := Ae(GA )−2e(H) (−A2 − A−2 )f (H) is the contribution of a spanning sub-graph H to D . Let the first sth-coefficient of D from M (D) be the coefficient of AM (D)−4(s−1) in D . A spanning sub-graph H contributes to the first sth coefficient of D if and only if v(H) − k(H) + g(H) ≤ s − 1, see [24, Theorem n (q) from 6.1]. This is used in [24] to give a formula of the penultimate coefficient of JK hn (D) when K has an A-adequate diagram. Theorem 2.2.3 can be rephrased as saying that the first n − 1 coefficients of Dn from M (Dn ) are equal to zero. Our proof will depend heavily on Theorem 2.2.6 and the restriction on v(H) − k(H) for a spanning sub-graph H that contributes to the first n − 1 coefficients of Dn . Definition 2.2.7. Let H ⊆ GA be a spanning sub-graph and σ be the Kauffman state that chooses the B-resolution on a crossing corresponding to an edge in H and the A-resolution on a crossing corresponding to an edge not in H. We associate to H the state graph H of σ. The following important lemma counts the faces of a spanning sub-graph by the number of state circles in its associate state graph. Lemma 2.2.8. The number of state circles in H is equal to the number of faces of H. Proof. Let σ be a Kauffman state of a link diagram D and σ ˆ be its dual state which resolves each crossing of the link the opposite way from σ. This means that if σ chooses the Aresolution at one crossing, then σ ˆ chooses the B-resolution at that crossing, and vice versa. 24 The lemma follows from [23, Lemma 4.2], which says that f (Gσ ) = |sσˆ D|, where Gσ is a ribbon graph constructed from σ. A crossing of a link diagram is nugatory if there is a closed curve in the projection plane meeting the diagram transversely at that crossing and does not intersect the diagram anywhere else. A link diagram is reduced if it contains no nugatory crossings. Throughout this section, D denotes a reduced, non A-adequate link diagram, Dn is its n-cable, and Gn A is the ribbon graph constructed as in Section 2.2 from the all-A state of Dn . A spanning n sub-graph H ⊆ Gn A is obtained from GA by deleting edges, with the associated state graph n associated H as in Definition 2.2.7. We will mainly be interested in the all-A state graph HA to the spanning sub-graph of no edges. By Theorem 2.2.6, we have Dn = e(Gn A )−2e(H) H⊆Gn A A (−A2 − A−2 )f (H)−1 , where H ranges through all spanning sub-graphs of Gn A . Recall that e(Gn A )−2e(H) XH = A (−A2 − A−2 )f (H)−1 is the contribution of H to Dn . We will first prove a lemma that allow us to sum the terms XH in a certain way to show the cancellation of the first n − 1 coefficients of Dn . 25 2.3 A cancellation lemma Lemma 2.3.1. Let c, d be integers and δ = (−A2 − A−2 ), the maximum degree of the sum k k Ac−2i δ d+i i i=0 in A is less than or equal to c + 2d − 4k. Proof. We factor out Ac δ d to obtain Ac δ d k k A−2i δ i . i i=0 Note first that A−2i δ i has the same degree 0 for each i, and each product A−2i δ i only has non-zero coefficients for terms of degree A−4j for j a nonnegative integer. For j > 0, the jth coefficient of the polynomial Ac δ d k k i=0 i A−2i δ i of the term Ac+2d−4(j−1) is given by the following equation. k The jth coefficient = (−1)i i≥j−1 26 k i i . j−1 To prove the lemma, it suffices to show that the jth coefficient is zero for all j ≤ k. Let p(x) be the polynomial x(x − 1)(x − 2) · · · (x − j + 2). This is a polynomial of degree less than k. The sum on the right for the jth coefficient becomes k (−1)i i=0 k p(i) . i (j − 1)! In order to show that this is zero, consider the binomial expansion of (1 + x)k : (1 + x)k k = i=0 k i x. i Taking the derivative with respect to x (j − 1)-times on both sides yields k · (k − 1) · · · (k − j + 2)(1 + x)k−j+1 k = i=0 k i(i − 1) · · · (i − j + 2)xi−j+1 , i where we have p(i) on the right side multiplying xi−j+1 . Setting x = −1 gives k 0= i=0 (−1)i−j+1 k p(i). i Then we just multiply by (−1)j−1 /(j − 1)! to obtain the lemma. 2.4 Labeling edges in HAn . By the remark immediately following Theorem 2.2.6, XH contributes to the first n − 1 coefficients of Dn if and only if v(H) − k(H) ≤ n − 2. Therefore, we may restrict to this 27 set of spanning sub-graphs and show that   d∗  XH  ≤ M (Dn ) − 4(n − 1), v(H)−k(H)≤n−2 to prove Theorem 2.2.3. Our strategy is to divide up the sum over equivalence classes where Lemma 2.3.1 can be repeatedly applied. In order to define this equivalence relation, we introduce labels and define sequences for each H that satisfies v(H) − k(H) ≤ n − 2 in this section. Since D is not A-adequate, there is an edge in the all-A state graph of D with two ends on the same state circle S, such that it lies in a region bounded by S on the projection 2-sphere. We name the crossing in D corresponding to this edge by e, and denote the n2 crossings corresponding to the cable of e in Dn by en . The set of crossings in en corresponding to n n will be denoted by en . loops in the all-A state graph HA See Figure 2.3 for the local picture of the resulting state graph where we choose the A-resolution at all the crossings in en . Figure 2.3: Identical local picture of the A-resolution of the 3-cable of a crossing. n is the all-A state graph of D n associated to the spanning sub-graph of Recall that HA n Gn A with no edges. We detail below labeling conventions for the edges corresponding to e n , and we will just refer to these edges by en . For an illustration of the labels, where in HA 28 we locally straighten the strands, see Figure 2.4. n there will be n state circles corresponding to the cabling of S in D n . We denote (a) In HA the state circle where the two ends of edges in en are by S0 . Then, Si is the state circle where one of its complementary regions on S 2 contains only en and S0 , . . . , Si−1 . The state circles S0 , S1 , . . . , Sn−1 divide the sphere into n + 1 regions. We denote the region with boundary Si−1 and Si for i = 1, 2, . . . , n − 1 by Ωi and the region with boundary S0 containing en by Ω0 . With this labeling we can refer to edges in en by which regions they are in. (b) For each i, an edge in Ωi corresponding to a crossing in en will be labeled by odd or even integers. First we label the edges in en by consecutive odd integers 1, 3, 5, . . . 2n − 1. For i = 0, an edge in Ωi with an end on Si−1 and an end on Si is labeled with the integer m if the end of the edge on Si−1 is between the ends of two edges labeled m − 1 and m + 1. (c) Drawing a curve through all the edges in Ω0 and dividing the projection sphere in half, an edge of en in Ωi for i ≥ 1 will be labeled L or R depending on whether it is on one side of the curve or the other side, respectively, so LΩi and RΩi refer to all the edges in en in region Ωi on the left side or on the right side, and we just let LΩ0 = RΩ0 = Ω0 . This notation, along with the numbering of edges of en , specifies each edge in en uniquely. n , so that for The crossings in en inherit the labels from the corresponding edges in HA any state graph of Dn , a segment in the state graph resulting from choosing a resolution at a crossing of en will also have the same label as the crossing. The numbers labeling en 29 Ω2 Ω2 Ω1 Ω1 Ω0 Ω0 Ω1 Ω1 Ω2 Ω2 3 2 1 4 3 2 RΩ1 5 4 3 3 RΩ2 Ω0 LΩ1 LΩ2 2 1 4 3 2 5 4 3 Figure 2.4: From left to right: an illustration of the labeling on the edges of en when n = 3. induce an ordering as well. We say that an edge with label a is smaller (resp. greater ) than an edge with label b if a < b (resp. b > a). Two edges are equal if their labeling numbers are equal. An edge with label c is between edges with label a and b if a < c < b. As long as it is clear which region Ωi we are referring to, we will refer to the edges by their labeling numbers and by which side they are on. For example, La ∈ Ωi and a ∈ LΩi for a an integer both indicate the edge in LΩi with label a. n uniquely corresponds to an edge of Gn . We say that an edge in H n is Each edge in HA A A n included in a spanning sub-graph H ⊆ Gn A , if H includes the corresponding edge in GA . We n are included in H. Schematically, the edges represent H by indicating which edges of HA n . This representation will allow us to compare that are included will be shown in red on HA n . We do this the number of state circles in H, the state graph associated to H, to that of HA n locally at each crossing which corresponds to by modifying the state circles sA (Dn ) of HA an edge included in H, replacing the A-resolution by the B-resolution. This will be crucial n is modified to proving the lemmas in this section. See Figure 2.5 and 2.6 for exactly how HA to a different state graph when one changes from the A-resolution to the B-resolution at a crossing. Recall that our goal is to organize the spanning sub-graphs H with v(H) − k(H) ≤ n − 2 into equivalence classes where Lemma 2.3.1 can be applied. In [69], this was done for spanning 30 B A Figure 2.5: Locally, we replace the two segments making up the state circle (in red) and the segment recording the crossing location (dotted) of the A-resolution by that of the segments coming from the B-resolution. Figure 2.6: From left to right: an all A-state graph with 3 state circles; an edge inclusion representing a spanning sub-graph of the ribbon graph constructed from the all A-state graph; the state graph associated to the spanning sub-graph, which has 2 state circles. sub-graphs H with v(H) − k(H) ≤ 1. Two spanning sub-graphs are in the same equivalence class if they only differ by what they include from en . The inclusion of each edge from en increases the number of faces by one, since en are one-edge loops for all of their associated state graphs. As a result, we set up sums of the form en e(Gn A )−2i A (−A2 − A−2 )f (H0 )+i−1 , i=0 where H0 is the member of the equivalence class that does not include any edges from en , and Lemma 2.3.1 can then be applied. When v(H) − k(H) > 1, the situation is more complicated. We can no longer set up equivalence classes based on edge inclusions in en , because it is no longer true that inclusion of an edge in en increases the number of faces by one. Instead, we generalize this specific example to defining a set of edges G(H) for a spanning sub-graph H, where the increase in 31 the number of faces is the same as the number of edges included. In order to control the change in the number of faces from edge inclusions, we make use of the explicit labelings just introduced, and we use Lemma 2.2.8 to compute f (H) from the state graph H. We first define a(H), which determines the set of edges in en we can add to H in any combination while splitting off as many state circles as the edges added. Definition 2.4.1. Given a spanning sub-graph H ⊂ Gn A , we define H to be the spanning sub-graph obtained from H by deleting from H all the edges outside of the region Ω1 . Consider the associated state graph H . Let m ∈ en be the smallest edge that has two ends on distinct state circles in H , and M ∈ en be the largest. We define a(H) to be the subset of edges c ∈ en such that m ≤ c ≤ M , and the empty set if every edge in en has two ends on the same state circle in H . For example, a(H) is empty for the spanning sub-graph with no edges and sub-graphs for which only a single edge in Ω1 is included. If a(H) is empty, then we can group them with other spanning sub-graphs where the only difference is which edges they include from en and apply Lemma 2.3.1. For spanning sub-graphs where a(H) is non-empty, we define sequences which controls the quantity v(H)− k(H) + g(H). The following lemma allows us to define the first terms of the sequences. Lemma 2.4.2. Suppose a(H) is not empty and v(H) − k(H) ≤ n − 2. Let m, M be the smallest and largest edge in a(H), respectively. (a) If 1 ∈ / a(H), at least one of the pair of edges R(m − 1), L(m − 1) in Ω1 is included in H. If 2n − 1 ∈ / a(H), at least one of the pair of edges R(M + 1), L(M + 1) is included in H. If 2n − 1 ∈ a(H), then an edge in Ω1 with an end attached to the portion of Ω0 cut off by 2n − 1 is included in H. See Figure 2.7 for an illustration. 32 (b) If 1 ∈ a(H): Let He be the sub-graph obtained from H by deleting all the edges in en and in the region Ωn−1 . There is an integer 0 ≤ k ≤ n − 2 such that the edge with the smallest labeling number in RΩk has two ends on the same state circle in the state graph H e associated to He . Proof. (a) The proof for the cases 1 ∈ / a(H), 2n − 1 ∈ / a(H), or 2n − 1 ∈ a(H) are similar, and we treat them together here. This follows from the fact that the edges m and M need to be attached to separate state circles in H , which has a simple picture, see Figure 2.7, only involving modifying n based on edge inclusions of H in Ω . Requisite edges of the state circles S0 , S1 of HA 1 Ω1 need to be included in H for m, M to have two ends on distinct state circles in the associated state graph. 1 m M 2n − 1 Figure 2.7: On the left: At least one of the two edges smaller than m in Ω1 needs to be included in order that m ends up with two ends on distinct state circles in H . The same is true with the edges bigger than M . On the right, the portion of the region Ω0 cut off by the edge 2n − 1 is shown in yellow, with the edges with one end attached to it. (b) For the case 1 ∈ a(H). We show that there exists an index 0 ≤ i ≤ n − 2 such that the edge with label i + 1 in RΩi has two ends on the same state circle in H e . For each edge labeled i + 1 in RΩi , 33 we let the pieces of state circles where the two ends of the edge are on in H e inherit n , so the top and the bottom pieces will be denoted as S the same labeling as in HA i and Si−1 , respectively. If the two boundaries are joined to each other in H e , then the edge i + 1 will have two ends on the same state circle in H e . We show the following statement by induction on i, assuming that a(H) is non-empty and contains the edge labeled 1. ◦ Let 0 < i ≤ n − 2. If for all 0 ≤ r ≤ i, the edge labeled r in RΩr has two ends on distinct state circles in He , then two edges ei ∈ Ωi and ei+1 ∈ Ωi+1 that are not in en , are included in He . Moreover, when we traverse the piece of the state n in the clockwise direction, we meet an end of e before we meet circle Si in HA i an end of ei+1 , and no edges between ei and ei+1 in Ωi+1 are included in He . This statement implies the lemma for the case 1 ∈ a(H) since, if the k as in the lemma does not exist, then the index n − 3 satisfies the criteria for the statement. We then have that there is an edge en−3 and en−2 as in the statement, but in He no edge is included in Ωn−1 . In RΩn−2 , the edge with label n − 2 will have two ends on the same state circle since Sn−3 would have no where to exit except to join back with Sn−2 . This is a contradiction. For i = 1, if the edge labeled 1 in Ω0 has two ends on distinct state circles in H , an edge e1 ∈ Ω1 that is not in en must be included in H. Otherwise, the boundaries on S0 where the two ends of edge 1 lie will join together and become the boundary of a single state circle in H . If now the edge labeled 2 in RΩ1 also has two ends on distinct state circles in He , we show that an edge e2 ∈ RΩ2 fitting the description of the statement must also be included in He . Since applying a Kauffman state to obtain 34 a state graph gives disjoint state circles, the boundary of S1 in H e can only exit the region enclosed by the state circle consisting of S0 by having He include an edge e2 in Ω2 . If we traverse S1 in the clockwise direction we would meet an end of e1 before we meet an end of e2 . Take the closest e2 to e1 in the counter-clockwise direction to be e2 . This proves the base case. See Figure 2.8 for an illustration where we straighten the strands locally. Clockwise direction RΩ2 RΩ1 4 e2 3 e1 2 1 Ω0 5 4 6 3 5 Figure 2.8: The base case i = 1. Now for the induction step: Assume that the statement is true for i − 1. We already have edges e1 , . . . , ei , each successive er , er+1 fitting the description of the statement by the induction hypothesis. So we only need to show that ei+1 as described in the statement exists. The state circle with boundary Si−1 will enclose a region in H e where the boundary Si must exit through an edge ei+1 ∈ Ωi+1 to avoid joining with Si+1 . Thus, an edge in RΩi+1 must be included in H e . Again we pick the one furtherest in the counter-clockwise direction to be ei+1 . This completes the proof of the statement. n, Definition 2.4.3. For each H where a(H) is non-empty, we define sa , an edge in HA according to the following rule: 35 • If a(H) contains 1: Pick the smallest index k ≥ 0 for which the edge labeled k + 1 in RΩk has two edge on the same state circle in He . This exists by Lemma 2.4.2. Let sa be the edge labeled k + 1 in RΩk if the edge labeled k in RΩk−1 is not included in H, and let sa be the edge labeled k in RΩk−1 otherwise. • If a(H) does not contain 1: Let m be the edge with the smallest label in a(H), we define sa :=      R(m − 1) in Ω1 , if R(m − 1) in Ω1 is included in H.     L(m − 1) in Ω1 , if R(m − 1) in Ω1 is not included in H. The edge sa provides us with a starting point to find an edge with two ends on the same state circle in the state graph associated to H, that is not included in H originally. If such an edge is included in H, the spanning sub-graph H resulting from this inclusion will have f (H ) = f (H)+1, and v(H )−k(H )+g(H ) = v(H)−k(H)+g(H) as desired. The sequences {tj }, {bj } to be defined below will allow us to keep track of any edge-inclusions of H that may result in an edge that has ends on distinct state circles in the state graph associated to H. Definition 2.4.4. Consider a spanning sub-graph H with sa as in Definition 2.4.3. Let i = 0, 1, 2, . . . n − 1, we will define for each i and each side L, R, sequences of edges {tj } and n . Recall that m, M are the smallest and largest labels of the edges {bj } in LΩi , RΩi in HA in a(H). See Figure 2.9 for an example. 36 • If sa is in RΩk , we define the sequences {tj }, {bj } for RΩi first.      sa      t0 = b0 := the smallest edge in RΩ greater than i         the smallest edge in RΩi greater than if i = k b0 of RΩi−1 if i > k b0 of RΩi+1 if i < k. • We set the first term of the sequences of LΩ0 equal to the first term of the sequences of RΩ0 . For LΩi , the first terms of the sequences are given by t0 = b0 := the smallest edge in LΩi greater than b0 of LΩi−1 . • If sa is in LΩk , we replace R by L and L by R in the above definitions and we get the definition for the first terms of the sequences for this case. • The rest of the sequences is given recursively by: For t0 ∈ RΩi (resp. LΩi ), tj := the smallest edge ≤ M in RΩi+j (mod 2) (resp. LΩi+(j mod 2) ) included in H greater than tj−1 . If there is no such edge then it is empty. For b0 ∈ RΩi (resp. LΩi ), bj := the smallest edge ≤ M in RΩi−j (mod 2) (resp. LΩi−j (mod 2) ) included in H greater than bj−1 . If there is no such edge then it is empty. 37 When i = 0, only the sequences {tj } are defined for RΩ0 and LΩ0 by the definitions above. One has terms from RΩ1 and another has terms from LΩ1 , respectively. We let {bj } for RΩ0 be the same sequence as {tj } of LΩ0 and {bj } for LΩ0 be the same sequence as {tj } of RΩ0 . 3 RΩ2 RΩ1 Ω0 2 1 5 RΩ1 RΩ1 3 RΩ2 4 3 3 RΩ2 2 4 Ω0 2 1 4 3 RΩ1 5 Ω0 2 1 4 3 5 Figure 2.9: Let sa = 2 ∈ RΩ1 for this example. The point of the sequences is to rule out edges in H that may make edges after the edge labeled 2 in RΩ1 have two ends on distinct state circles in H. The sequence for RΩ1 are {bj } = {2, 5, ∅, . . .} and {tj } = {2, ∅, . . .}. Definition 2.4.5. Let the sequences {tj } and {bj } for LΩi or RΩi be given, and let 0 ≤ n. i ≤ n − 2. We define subsets ti and bi which we use to define G(H), a subset of edges in HA Recall that an edge with label b is between edges labeled a and c if a < b < c. See Figure 2.10 for what kind of edges this picks out relative to the sequences. For t0 ∈ RΩi (resp. LΩi ), • ti := union of all edges of en in RΩi (resp. LΩi ) between tj and tj+1 for j even and tj j non-empty. If tj+1 is empty then we take all edges after tRi and ≤ M in RΩi (resp. LΩi ). • bi := union of all edges of en in RΩi (resp. LΩi ) between bj and bj+1 for j even and bj j non-empty. If bj+1 is empty then we take all edges after bRi and ≤ M in RΩi (resp. LΩi ). Finally, we define G(H) := ∪t0 ∈RΩ (ti ∩ bi ) ∪ ∪t0 ∈LΩ (ti ∩ bi ) ∪ sa i i 38 Figure 2.10: On the left: All the edges shown in blue here after the red edge are candidates to be included in G(H), as long as there are no edges above and below after the middle red edge that are included in H. In the middle and on the right: The inclusion of the top and bottom edges in red means that only three edges in blue after red will be included in G(H). This is so that they will still have two ends on the same state circle as shown on the right when we only resolve the crossings corresponding to the red edges. if 1 ∈ a(H), and G(H) := ∪t0 ∈RΩ (ti ∩ bi ) ∪ ∪t0 ∈LΩ (ti ∩ bi ) i i if 1 ∈ / a(H). Lastly, If a(H) = {∅}, then G(H) := {∅}. 2.5 Equivalence classes based on G(H). The point of the string of definitions and lemmas leading to the definition of a(H) and G(H) is that it gives us a way to understand how the number of faces changes for a new spanning sub-graph obtained by adding edges. We define an equivalence relation that will put spanning sub-graphs into equivalence classes where we can control the differences in the combinatorial quantities e(H) and f (H), directly related to v(H) − k(H) + g(H), from one member to another. Definition 2.5.1. We define an equivalence relation R on the set of spanning sub-graphs H ⊆ Gn A where v(H) − k(H) ≤ n − 2. We say that H∼ =R H 39 if 1. a(H) = a(H ). 2. G(H) = G(H ), and H only differs from H by what edge it includes from G(H) ∪ n. (en a(H)). Recall that en is the set of loops with both ends on S0 in HA This is an equivalence relation because a(H) and G(H) are well-defined on H. Within an equivalence class C of R, the members of C only differ from each other by edge inclusion in G(H) ∪ (en a(H)). Let H be a member of an equivalence class C and consider the spanning sub-graph H0 obtained from H by deleting all the edges in G(H)∪(en a(H)). This spanning sub-graph is well-defined for C since a(H) and G(H) are the same across C. Lemma 2.5.2. Let C be an equivalence class of the relation R, let H be a member of C and H0 be the spanning sub-graph obtained from H by deleting all the edges in G(H)∪(en a(H)). Then (a) H0 is also in C. (b) Adding any combination of k edges in G(H) ∪ (en a(H)) to H0 gives another member H of C. In addition, f (H ) = f (H0 ) + k and v(H ) − k(H ) + g(H ) = v(H0 ) − k(H0 ) + g(H0 ). Proof. For (a), we need to show that a(H0 ) = a(H) and G(H0 ) = G(H). Consider the spanning sub-graph H0 obtained from H0 by deleting all the edges in Ω0 and outside of Ω1 as in Definition 2.4.1. Let m and M be the smallest edge and the largest edge in a(H). The set a(H0 ) would only differ from a(H) if we delete edges in Ω1 that are not between m and 40 M . However, the algorithm for G(H) requires that all edges in G(H) are between m and M . Adding or deleting edges in G(H) ∪ (en a(H)) does not affect He , as used in Lemma 2.4.2 to define sa , and the sequences, since G(H) ∪ (en a(H)) ⊂ en , so G(H0 ) = G(H) and we have that H and H0 are in the same equivalence class. For (b), an edge is in G(H0 ) if and only if it is between a pair of edges e and e where e ∈ RΩi and e ∈ RΩi+1 or e ∈ RΩi−1 , and there are no edges in RΩi−1 or RΩi+1 included between e and e , or it is sa . We have the same condition for an edge in G(H0 ) on the left side by switching all the labels from R to L. See Figure 2.11 for an example. Figure 2.11: From left to right: An example of edges in G(H) marked in blue; the state graph resulting from including zero edges from G(H); the state graph resulting from choosing one edge from G(H), note that a state circle splits off; the state graph resulting from including two edges from G(H) with two state circles split off. By design and since we are beginning the sequences at sa which are especially chosen through using Lemma 2.4.2, all the edges in G(H0 ) ∪ (en a(H)) have two ends on the same state circle in state graph H0 associated to H0 , with the identical local picture as shown in Figure 2.11: they all have two ends on the same state circle in H0 , and none of the pairs of edges from G(H0 )∪(en a(H)) lies embedded on distinct sides of a state circle such that only one end of an edge lies between the two ends of another. Thus, including any combination of k edges to H0 splits off k additional state circles in the new spanning sub-graph H, and this shows the first part of (b). Since we cannot add an edge in region Ωi without there already being an edge included in Ωi from the definition of the sequences which define G(H), we also have that v(H )−k(H ) = 41 v(H0 ) − k(H0 ). This also shows that g(H ) = g(H0 ) since the increase in the number of faces is the same as the number of edges included. 2.6 Combinatorics of G(H) and ribbon graphs By Lemma 2.5.2, if C is an equivalence class of the relation R, and r = |G(H) ∪ (en a(H))| is the cardinality of C, we may write r H∈C XH = i=0 r e(Gn )−2e(H0 )−2r A A (−A2 − A−2 )f (H0 )+r−1 . i For each equivalence class C, we are almost ready to use Lemma 2.3.1 to show the cancellation of the top coefficients of this sum. However, we need an estimate on r so that we know how many coefficients are cancelled. We also need to know how the resulting upper bound on the maximum degree compares with M (Dn ). Our main result of this section is the following: Lemma 2.6.1. If a spanning sub-graph H ⊂ Gn A satisfies v(H) − k(H) + g(H) ≤ n − 2, then |G(H)| + v(H) − k(H) + g(H) ≥ |a(H)| − 1. This lemma connects the cardinality of G(H) to the combinatorial ribbon graph quantities v(H), k(H), and g(H) and is the last important piece of the proof of Theorem 2.2.3. The reader may skip ahead to Section 2.7 to see how it is applied and come back to this section later. To prove this lemma, we make use of the explicit labeling from Section 2.4. Lemma 2.6.2. Let e1 be an edge in RΩi included in a spanning sub-graph H. Suppose another spanning sub-graph H only differs from H by including two additional edges e2 in RΩi and e3 in RΩi+1 that satisfy the following criteria: 42 • The edges e2 and e3 are not included in H. • The edge e2 is between e1 and e3 . • None of the edges in RΩi and RΩi−1 between e1 and e3 are included in H, and none of the edges in RΩi+1 between e1 and e2 are included in H. Then v(H ) − k(H ) + g(H ) = v(H) − k(H) + g(H) + 1. The same statement for LΩi can be obtained by replacing all the labels from R to L. e2 e1 e1 e3 Figure 2.12: On the left we have H and on the right we have H . Their local pictures only differ in the inclusion of two edges e1 and e2 . Proof. An example satisfying the hypotheses of the lemma is shown in Figure 2.12. By the definition of g(H), 2v(H) − 2k(H) + 2k(H) − v(H) + e(H) − f (H) 2 v(H) + e(H) − f (H) = . 2 v(H) − k(H) + g(H) = Compare the state graph H and H . Including the edge e3 will first split off a state circle. Then, including e2 joins the state circle split off by e1 , e3 to another state circle. The end result is that the number of states circles of H and H are the same. By Lemma 2.2.8, this 43 means f (H ) = f (H). Therefore, v(H ) + e(H ) − f (H ) 2 v(H) + e(H) + 2 − f (H) = 2 v(H ) − k(H ) + g(H ) = = v(H) − k(H) + g(H) + 1. We define ga (H), which counts the occurrences of edge inclusions described in Lemma 2.6.2 from the sequences {tj }, {bj } in Definition 2.4.4. Definition 2.6.3. If sa as in Definition 2.4.3 is in RΩk , we define ga (H) by taking the union of tgi and bgi defined for LΩi and RΩi below: For RΩi (resp. LΩi ) and the sequences {tj }, {bj }. • For all i, tgi is the number of even ’s such that t is non-empty and between two non-empty terms bj and bj+1 where j is even. • For i = k, bgi is the number of even ’s such that b is non-empty and between two non-empty terms tj and tj+1 where j is even. • bgk is #{bj } 2 where #{bj } is the number of non-empty terms in the sequence {bj }, and · is the floor function which rounds down to the nearest integer. Lastly, ga (H) := bgi + 0≤i≤k for each RΩi tgi + k≤i≤n−2 for each RΩi 44 tgi . 1≤i≤n−2 for each LΩi If sa is on the left side, we replace L by R and R by L in the above definition to get the definition of ga (H) in this case. By Lemma 2.6.2, ga (H) counts the quantity v(H) − k(H) + g(H) locally. Now we are ready to show Lemma 2.6.1. Proof. We assume first that sa ∈ RΩk and M is the largest label of the edges in a(H). The case where sa ∈ LΩk is the same by reflection. Recall that G(H) is defined in Definition 2.4.5 by taking the union of the intersection bi ∩ ti for all i ranging over RΩi and LΩi . The cardinality of the intersection bi ∩ ti is related to tgi , bgi , bi+1 , and ti−1 when i ≤ n − 2 for RΩi (resp. LΩi ) as follows: |bi ∩ ti | = |bi | − |tgi | − |bi | (2.2) |bi ∩ ti | = |ti | − |bgi | − |ti |, where bi consists of edges in RΩi between bj and bj+1 in the sequence {bj } for RΩi+1 where j is even and bj is non-empty. Similarly, ti consists of edges in RΩi between tj and tj+1 in the sequence {tj } for RΩi−1 where j is even and tj is non-empty. These are the edges excluded from bi and ti by the intersection bi ∩ ti , respectively. We also have |bi+1 | ≥ |bi | or |bi+1 | = |bi | − 1. (2.3) |ti−1 | ≥ |ti | or |ti | = |ti−1 | − 1. By Definition 2.4.5, an edge ≤ M of RΩi (resp. LΩi ) with sequence {bj } and {tj } is in bi , if and only if it is between bj and bj+1 where j is even and bj is non-empty. An edge is 45 excluded from bi by the intersection bi ∩ ti if and only if it is between tj and tj+1 where j is odd and bj is non-empty. The sequence {bj } for RΩi+1 is the sequence {tj } for RΩi with the first term removed. Therefore, bi+1 contains edges excluded from bi with one added to all the labels. Since RΩi+1 has one less edge than RΩi , we have the two possibilities for |bi+1 | and |bi |. Similarly, we have the statement for |ti | and |ti−1 | and for LΩi . We define the following quantities. • C := #{i : k ≤ i < n − 2, |bi+1 | = |bi | − 1} for each RΩi such that the edge lost from bi is an edge coming from bRk+1 in the following sense: the edge - r in RΩr is in br for each k + 1 ≤ r ≤ i. • C := #{i : 0 ≤ i < n − 2, |bi+1 | = |bi | − 1}, for each LΩi such that the edge lost from bi is an edge coming from tRk−1 in the following sense: the edge - r is in br for each 0 ≤ r ≤ i of LΩr and in tr for each 0 ≤ r ≤ k − 1 of RΩr . • E := #{i : k < i < n − 2, i is counted by C and i + k is counted by C}. As discussed, an edge in bi+1 is an edge in bi with one added to the label, so if |bi+1 | = |bi |−1 is counted by C, an edge with label n − i is included in bi . Similarly, an edge with label n−i−1 is included in bi−1 , and so on. We see that an edge with label R(n−i−(i−k −1)) = R(n − 2i + k + 1) is included in bk+1 . Similarly, an index i counted by C corresponds to an edge with label R(n−(i−k)−(k +i−k −1)) = R(n−2i+k +1) included in tk−1 . Therefore, E counts the number of edges with the same labeling numbers in bRk+1 and tRk−1 . To complete our estimate of |G(H)|, we show |bk ∩ tk | + |bk+1 | + |tk−1 | ≥ |sa | − |tgk | − |bgk | + δk + E, 46 (2.4) where |sa | is defined to be the number of edges greater than sa and ≤ M in RΩk , and δk is a quantity counted by C. To begin with, we have |bk | = |sa | − |bgk | − |tk−1 | since edges of the sequence {bj } for RΩk counted by bgk are not included in bk , and there are additional edges excluded from |sa | that are replaced by edges in tk−1 . Then, |bk ∩ tk | + |bk+1 | ≥ |sa | − |bgk | − |tk−1 | − |tgk | − δk + E, where δk = 1 if |bk+1 | = |bk | − 1 and 0 otherwise. So, δk is counted by C. The edges greater than sa that are part of the sequence {tj } of RΩk+1 , counted by tgk , will not be included in bk ∩ tk . In addition, extra edges with the same labels as those in tk−1 may be included in bk+1 . These are counted by E. Adding |tk−1 | to the left hand side of the inequality produces the inequality (2.4). Putting the statements (2.2) and (2.3) together for each 0 ≤ i ≤ n−2 and using the definition of G(H), we have |G(H)| ≥ |bk ∩ tk | + |bi ∩ ti | + i=k for each RΩi |bi ∩ ti | i=0 for each LΩi ≥ |bk ∩ tk | + |bk+1 | + |tk−1 | −C −C − bgi − 0≤i k where i ≤ the highest index counted by C. 48 (b) The second set are the rest of the edges in all the sequences {tj }, {bj } defined for H. 1 We denote the spanning sub-graph obtained by adding the set (a) to Hn A by H , and the spanning sub-graph obtained by adding the set (b) to H1 by H2 . The spanning sub-graph H is obtained by adding the rest of the edges included in H to H2 . We first show that v(H1 ) − k(H1 ) ≥ C + k − 1. (2.9) In H1 , the edges from V are edges between state circles Si , Si+1 for 0 ≤ i ≤ k − 2 in n , the all-A state graph associated to Hn , which increases as i increases. Adding each HA A edge counted by C also increases v(H1 ) − k(H1 ) by 1, since it merges different state circles from previous edge-inclusions, see Figure 2.13. RΩk RΩ1 Figure 2.13: The edgeset V is shown in black. Adding the set (b) of edges from the rest of the sequences to get H2 , we apply Lemma 2.6.2 repeatedly. The increase in g(H) is given by ga (H) since we only counted the increase to v(H1 ) − k(H1 ) + g(H1 ) from increasing sequence of edges in set (a). We get the inequality (2.7). For (2.8), we consider the set of edges e in RΩk+1 which do not belong to bk+1 ∩ tk+1 . This is a set with cardinality ≥ C − E. Let s be the highest index of RΩs counted by C. By definition, the edge cs lost from bi comes from bRk+1 . Let cr be the edge corresponding 49 to cs − r in RΩs+1−r . Now we have a set of increasing edges all less than cs − r for each r in H2 . We also have |bk+1 | ≥ |e | + |bk |, or |bk+2 | ≥ |e | + |bk+1 |, since otherwise, cr would not belong to br for some r, see Figure 2.14. RΩk+2 RΩk+1 RΩk RΩk−1 Figure 2.14: The leftmost figure depicts a possible set of bk shown in blue, and the edge excluded from bk by the edge in RΩk−1 . Depending on the sequences {tj }, {bj } for RΩk and RΩk+1 , we have at least an extra edge back from bk+1 or bk+2 . Combing this with (2.5), we note that any additional loss of edges from e as we increase r increases one of v(H2 ) − k(H2 ), g(H2 ), or G(H2 ) by 1 independent of (2.9) and ga (H). So v(H) − k(H) + g(H) + |G(H)| ≥ |sa | + E − (C + C ) + C + k − 1 + C − E, and we are done. 50 2.7 Proof of Theorem 2.2.3 To summarize, by Theorem 2.2.6, we have that Dn = H⊆Gn A Ae(GA )−2e(H) (−A2 − A−2 )f (H) . n is given by The contribution of a spanning sub-graph H ⊂ Gn A to D e(Gn A )−2e(H) XH := A (−A2 − A−2 )f (H) . Analysis of the monomials involved in the expansion (see [24], Theorem 6.1) show that a spanning sub-graph H contributes to the first n − 1 coefficients if and only if v(H) − k(H) + g(H) ≤ n − 2. We organize the contributions of these spanning sub-graphs by putting them into equivalence classes defined by R as in Definition 2.5.1. Let C denote an equivalence class of R. Recall that d∗ of a polynomial p(A) in A is the maximum degree of p(A). We have d∗ D n = d∗ C XC , where XC := H∈C and C ranges over equivalence classes of R. 51 XH , Let r = |G(H0 ) ∪ (en a(H0 ))|, where H0 is the member in C which does not include any edges in G(H0 ). By Lemma 2.6.1, r ≥ n − 1 − (v(H0 ) − k(H0 ) + g(H0 )). We also have, by Lemma 2.5.2, r XC = i=0 r e(Gn )−2e(H0 )−2r A A (−A2 − A−2 )f (H0 )+r . i Applying Lemma 2.3.1 to this sum and writing f (H0 ) in terms of v(H0 ), g(H0 ), and k(H0 ) using the definition of genus, we get d∗ XC ≤ c(Dn ) − 2e(H0 ) + 2(−2g(H0 ) + 2k(H0 ) − v(H0 ) + e(H0 )) − 4r ≤ c(Dn ) − 2e(H0 ) + 2(−2g(H0 ) + 2k(H0 ) − v(H0 ) + e(H0 )) − 4(n − 1 − (v(H0 ) − k(H0 ) + g(H0 ))) ≤ c(Dn ) + 2v(H0 ) − 4(n − 1) ≤ n2 c(D) + 2|sA (Dn )| − 4(n − 1), where n2 c(D) + 2|sA (Dn )| = M (Dn ). Since degXC ≤ M (Dn ) − 4(n − 1) for all the equivalence classes C containing all spanning sub-graphs H with v(H) − k(H) ≤ n − 2, we see that d∗ Dn ≤ M (Dn ) − 4(n − 1). 2.8 A worked out example The proof of Theorem 2.2.3 uses the labeling on cables of a non A-adequate link diagram to separate the contributing spanning sub-graphs into equivalence classes. There is then cancellation of coefficients summing over the contributing polynomial of each member of the 52 equivalence class. In this section, we illustrate this process on an example. Figure 2.15: From left to right: a non A-adequate link diagram, the all-A state graph, and the all A-state graph of the 3-cable. The link diagram shown on the left in Figure 2.15 is not A-adequate because there is an edge with both ends on the same state circle S in the all-A state graph of the diagram. This link may have an A-adequate diagram. However, our aim is to show the division of spanning sub-graphs into equivalence classes for this non A-adequate diagram. We consider the case n = 3 and apply the construction for the proof of Theorem 2.2.3. The theorem in this case says that M (D3 ) − d∗ D3 ≥ 4 · (3 − 1). The first step is to restrict to spanning sub-graphs H ⊂ Gn A with v(H) − k(H) ≤ 1 by the remark following Theorem 2.2.6. Following the conventions of Section 2.2, e3 denotes the edges in the all A-state graph of D3 with two ends on a state circle S0 , S1 is the state circle coming from cabling S in D3 that contains S0 and no other state circles coming from cabling S in one of its complementary regions on S 2 . The edges in the all-A state graph of D3 corresponding to e3 have their labelings. We also label all the regions Ω0 and Ω1 . See Figure 2.16. The restriction v(H)−k(H) ≤ 1 means that only edges between a pair of distinct vertices, and one-edge loops e3 in Gn A , may be included in H. There are two cases, one where a(H) is empty and another where a(H) is non-empty. The case where a(H) is empty includes the case where the pair of distinct vertices with an edge included between them is not S0 53 and S1 , v(H) − k(H) = 0, and when a single edge or two edges with the same label in Ω1 are included between S0 and S1 . The spanning sub-graph H of each of these cases belongs to an equivalence class where the only difference between the members are which edges are included from e3 . Since |e3 | = 3, applying Lemma 2.3.1 will show that the contribution of H to terms with power greater than M (D3 ) − 8 in D3 is equal to zero. RΩ1 Ω0 2 4 1 3 5 LΩ1 2 4 Figure 2.16 We restrict to the local picture of Gn A involving only S0 and S1 . The case when a(H) is non-empty is what motivates the technicalities in the proof of Theorem 2.2.3. We cannot put H in an equivalence class where the only difference between the members is what edges it includes from e3 , since an edge in e3 may belong to a(H) where it has two edges on distinct state circles in the state graph associated to H. Inclusion of the edge may then give a spanning sub-graph H with a different V (H ) − k(H ) + g(H ), which determines the power of the term (−A2 − A−2 ) of the contribution XH . Lemma 2.3.1 will not apply to the sum of the contributions in this case. On the other hand, if a(H) is non-empty, then at least one edge must be included in H between S0 and S1 and not between any other pairs of vertices which makes our restriction to the local picture valid. If only one edge between S0 and S1 is included then a(H) is empty and we have already addressed this case. When more than one edge between S0 and S1 is included in H, we look to the edges in region Ω1 that have two ends on the same state circle 54 in the state graph associated to H to set up a equivalence class where we can apply Lemma 2.3.1. This is the primary motivation for our algorithm in the proof of Theorem 2.2.3. The general case is more complicated, but we can list all the possibilities for a(H) for the specific case we have here where n = 3. When a(H) is non-empty, the possible subsets are a(H) = {1}, {3}, {5}, {1, 3}, {1, 5}, {3, 5}, and {1, 3, 5}. In Table 2.3 we list all the possible sequences {bj }, {tj } and the corresponding G(H), v(H)−k(H), and g(H) for a(H) = {∅}, {1}, {3}, {5}, {1, 3}, {1, 5}, {3, 5}. 55 en a(H) {bj } for RΩ1 {tj } for RΩ0 {tj } for LΩ0 {bj } for LΩ1 {∅} {1, 3, 5} {∅} {∅} {∅} {∅} {1} {3, 5} {3} {1, 5} {5} {1, 3} {∅} {∅} {R2} {R2, 3} {R4} {R4, 5} {R2} {R2, 3} {∅} {1} {∅} {3} {∅} {5} {∅} {3} {∅} {1} {R2} {1, R2} {R2, 3} {1, R2, 3} {R2} {∅} {R2, 3} {3} {R2, 5} {5} {R2, 3, R4} {3, R4} {R2, 3, R4, 5} {3, R4, 5} {∅} {1} {∅} {3} {∅} {5} {∅} {3} {1} {1, L2} {1, L2, 3} {1} {1, L2} {1} {1, L2, 3} {∅} {3} {3, L4} {3, L4, 5} {5} {3} {3, L4} {3} {3, L4, 5} {∅} {∅} {∅} {∅} {∅} {∅} {∅} {∅} {∅} {L2} {L2, 3} {∅} {L2} {∅} {L2, 3} {∅} {∅} {L4} {L4, 5} {∅} {∅} {∅} {∅} {∅} {1, 3} {3, 5} {5} {1} 56 a(H) G(H) v(H) − k(H) 0 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {R2} 1 {R2} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 {R4} 1 {5} 1 {∅} 1 {∅} 1 {R4} 1 {∅} 1 {∅} 1 {∅} 1 {∅} 1 g(H) 0 0 0 1 0 1 0 1 0 1 1 1 2 1 1 2 2 0 1 1 2 1 1 1 1 2 Table 2.3: The tables detail the various possibilities for the equivalence classes corresponding to the different cases for a(H) and their corresponding values of G(H), v(H) − k(H), and g(H), when sa is in RΩ1 or Ω0 . The case for when sa is in LΩ1 is analogous. 56 For example, for a(H) to contain edges labeled 1 and 3, an edge attached to the portion of the state circle S0 cut off by 1 must be included in H, and one of the edges labeled 4 in LΩ1 and RΩ1 must be included. There will be various combinations for these edges that cut off state circles so that 1 ends up with two ends on different state circles in the state graph associated to H, but the only two cases that we need to deal with are • 1 is not included in H. • 1 is included in H. If 1 is not included in H, the edge with label R2 will have two ends on the same state circle in the state graph associated to H. This is the edge included in G(H). The sequence {bj } for RΩ1 will either be {R2} or {R2, 3} since we don’t include edges with labels larger than or equal to 4, and these are shown in the row starting with a(H) = {1, 3} in Table 2.3. On the other hand, if 1 is included, there is a corresponding increase in g(H) from 0 to 1. Taken altogether, v(H) − k(H) + g(H) ≥ 2 and so we do not need to account for the contribution of those equivalence classes. This is the idea behind Lemma 2.6.1, which estimates the cardinality of the set G(H) by v(H) − k(H) + g(H). The more terms there are in the sequences, and so the smaller |G(H)| is, the bigger v(H) − k(H) + g(H) will be, so that we will always end up with |G(H)| + v(H) − k(H) + g(H) ≥ a(H) − 1. As shown in Section 2.7, this controls the degree of the contribution of each equivalence class so that even if there is only a single element H in an equivalence class, deg XH ≤ M (Dn ) − 4(n − 1). 57 2.9 Detecting semi-adequacy using the colored Jones polynomial We use Theorem 1.1.1 to define a link invariant as in [69]. Let D be a diagram of an oriented link K. Let c− (D) be the number of negative crossings of D, and recall that |sA (D)| is the number of state circles in the all A-resolution of D, c(D) is the number of crossings in D, and w(D) is the writhe. We consider the complexity (c− (D), c(D), |sA (D)| − w(D)), ordered lexicographically. Let D(K) be the set of diagrams which minimizes this complexity. Let GD (n, A) be as defined in Section 1.2, see equation (1.2). We let M (GD (n, A)) := M (Dn−1 ) + 4 − 2(n − 1) − w(D)((n − 1)2 + 2(n − 1)). Recall that a lower bound hn (D) (See Definition 2.2.2.) of the minimum degree d(n) of JK (n, q) is − 41 M (GD (n, A)). Definition 2.9.1. Let K be a link and D an oriented link diagram in D(K). For i ≥ 1, let βi = βi (D) be the coefficient of AM (GD (i+2,A))−4(i−1) in GD (i + 2, A). Define A (q) JD ∞ := i=1 βi q i−1 . We will need the following lemma from [69]. Lemma 2.9.2. [69, Lemma 3.4] Suppose that for a link K, there is D ∈ D(K) that is 58 A-adequate. Then, all the diagrams in D(K) are A-adequate. Applying Theorem 1.1.1, we have the following corollaries. A (q) = 0 if and only if D is A-adequate. Corollary 2.9.3. JD Proof. By definition, the coefficient βi of AM (GD (i+2,A))−4(i−1) in GD (i + 2, A) is the coefficient of q hi+2 (D)+i−1 in JK (i + 2, q). If D is not A-adequate, then Theorem 1.1.1 says that d(i + 2) ≥ hi+2 (D) + i, so βi = 0 for all i. This shows the forward direction. For the converse, β1 is the coefficient of q h3 (D) in JK (3, q). If D is A-adequate, then h3 (D) is equal A (q) = 0. to the minimum degree d(3) of JK (3, q), so β1 = 0, and this shows that JD A (q) defined above is independent of the diagram D ∈ Corollary 2.9.4. The power series JD A (q). D(K), thus it is an invariant of K, which we denote by JK Proof. If K is not A-adequate, then any diagram in D(K) is not A-adequate. Let D be A (q) = 0. If K is A-adequate, then an Aa diagram in D(K), then by Theorem 1.1.1, JD adequate diagram of K minimizes the complexity (c− (D), c(D), |sA (D) − w(D)|), thus it belongs to D(K), and all the diagrams in D(K) are A-adequate by Lemma 2.9.2. Let D A (q) be a diagram in D(K). As shown in [9], the colored Jones polynomial has a tail and JD n (q)}∞ , therefore it is also independent records the stable coefficients of the sequence {JK n=2 of the diagram D. Definition 2.9.5. Consider the graph GA with vertices the state circles, and edges the segments from the graph of the A-resolution HA of D, we denote by χA (D) the Euler characteristic of GA . Corollary 2.9.6. Suppose D is an A-adequate diagram of a link K and D is another diagram of K. Then D is A-adequate if and only if c− (D) = c− (D ) and χA (D) = χA (D ). 59 Proof. If D is A-adequate, then c− (D) and |sA (D)|−w(D) are invariants of K [78, Theorem 5.13]. Thus, χA (D) = |sa (D)| − c(D) = |sA (D)| − w(D) − 2c− (D) is also an invariant of K. For the converse, since D is A-adequate, the minimum degree d(n + 1) of JK (n + 1, q) is equal to hn+1 (D), which we rewrite here slightly differently: 1 hn+1 (D) = − (n2 c(D) + 2n|sA (D)| − 2 + 4 − 2n − ω(D)(n2 + 2n)) 4 1 = − (2c− (D)n2 + 2n(|sA (D)| − w(D)) + 2 − 2n). 4 If D is not A-adequate and c− (D) = c− (D ), χA (D) = χA (D ), then sA (D) − ω(D) = sA (D ) − ω(D ), so hn+1 (D) = hn+1 (D ). Theorem 1.1.1 applied to D will imply that d(n + 1) < hn+1 (D), which is a contradiction. 60 Chapter 3 The Slope Conjecture and 3-string pretzel knots Outside the class of semi-adequate links, the Slope Conjecture has been verified for the class of 3-string pretzel knots P (−1/2, 1/3, 1/p) in the paper [41] by Garoufalidis where he introduced the Slope Conjecture and also verified it for alternating knots and knots with crossing number ≤ 9. In [28], Dunfield and Garoufalidis gave criteria for a spun-normal surface in an ideal triangulation to be incompressible. Based on this work, Garoufalidis and van der Veen proved the Slope Conjecture for 2-fusion knots [46]. Suppose that Kp,q is the (p, q)-cable of a knot for which the Slope Conjecture is true, Kalfagianni and Tran [70, 71] recently showed that the Slope Conjecture is also true for Kp,q provided certain conditions on the degrees are met. Using this, they proved the Slope Conjecture for iterated cables of torus knots and iterated torus knots. The work of this chapter is motivated by the following question. Question 3.1. How does the Jones slope “pick” the corresponding boundary slope for non semi-adequate links? To provide more examples to study Question 3.1, we prove: Theorem 1.1.2. Let r, s, t be nonzero integers greater than 5 such that r is even, s, t are odd, and r > s, t. The Slope Conjecture is true for pretzel knots P (−1/r, 1/s, 1/t). 61 We begin with the background information on essential surfaces in Section 3.1, whose boundary slopes may be enumerated by classifying branched surfaces. As we will see in Section 3.2, this can be applied to a 2-bridge knot to find all of its boundary slopes. This is the building block for the Hatcher-Oertel algorithm, which combines incompressible surfaces locally defined for each tangle into an incompressible surface for the Montesinos knot. We discuss how the surfaces are fitted together in Section 3.3 and describe the algorithm in Section 3.4. We will apply the algorithm to 3-string pretzel knots and compute the boundary slope of a specific incompressible surface. In Section 3.5, we describe the computation of the Jones slope for the pretzel knot, and show that it agrees with the boundary slope in Section 3.6. 3.1 Essential surfaces and branched surfaces For the rest of this chapter, we assume that M is a compact, orientable and irreducible 3-manifold with boundary ∂M and that all surfaces S ⊂ M are properly embedded, so ∂S ⊂ ∂M . A surface S is two-sided if its normal bundle in M is trivial. Definition 3.1.1. A properly embedded, two-sided, and connected surface S ⊂ M is incompressible if for each disk D ⊂ M with D ∩ S = ∂D there is a disk D ⊂ S with ∂D = ∂D. See Figure 3.1. The disk D is called a compressing disk for S if there is no D ⊂ S with ∂D = ∂D. S is ∂-incompressible if for each disk D ⊂ M such that ∂D is the union of two arcs α and β where α ⊂ S and β ⊂ ∂M , intersecting only at their endpoints, there is a disk D ⊂ S with boundary α ∪ β such that β ⊂ ∂S. The disk D is called a ∂-compressing disk for S if there is no D ⊂ S as above. See Figure 3.2. 62 D D D S S Figure 3.1: A compressing disk [52]. α α S D S D ∂M D ∂M Figure 3.2: A ∂-compressing disk [52]. Definition 3.1.2. A properly embedded surface S ⊂ M is essential if the boundary of its normal bundle in M is incompressible and ∂-incompressible. Remark 3.1.3. This definition of essential surface is consistent with the usual definition for two-sided surfaces, since the boundary of the normal bundle of S consists of two copies of S if S is two-sided. Definition 3.1.4. A branched surface B ⊂ M is a compact surface locally modeled on Figure 3.3a. i.e., B is locally homeomorphic to two planes R2 with the half-planes identified. A branched surface neighborhood of B, see Figure 3.3b, in M is a codimension 0, compact submanifold N (B) ⊂ M 3 foliated by interval fibers locally isomorphic to the model, where the branched surface is obtained by identifying the fibers to a point. A surface S ⊂ M is carried by B if S can be isotoped into the interior of N (B) so that it intersects the fibers transversely. Floyd and Oertel showed that there are only finitely many branched surfaces in M and 63 The proof of the theorem will follow fairly easily from a fundamental result of manifolds, which are aclosed subsets locally [FO] about branched The proof of thesurfaces theorem in will3follow fairly easily from fundamental result of diffeowhich are closed subsets locally diffeo[FO] about branched 3 manifolds, morphic to the modelsurfaces in the in first figure below. morphic to the model in the first figure below. (a) A branched surface. (b) A branched surface neighborhood. A branched surface B is said to carry a surface S if S lies in a fibered regular neighA branched surface B is to carry a and surface S if surface S liesneighborhood in a fibered regular neighFigure 3.3: of Models of asaid branched surface a branched [54]. borhood N(B) B , indicated in the second figure, and is transverse to all the fibers borhood of B ,all indicated second figure, is transverse toresult all the fibers of N(B)N(B) . If S meets the fibersin of the N(B) it is said to haveand positive weights. The that .itIfsuffices to enumerate branched for thesaid computation ofpositive boundaryweights. slopes of aThe result of of N(B) S meets all the fibers ofsurfaces N(B) it is to have M with incompressible boundary [FO] is that in a compact irreducible 3 manifold with boundary of [FO] isexist thatfinitely in a compact irreducible 3 manifold ∂M)incompressible such that the surthere many branched surfaces (Bi , ∂Bi ) ⊂M(M, knot [29]. are, ∂B exactly all the incompressible, faces carried with positive weights bysurfaces these Bi ’s(B there exist finitely many branched i i ) ⊂ (M, ∂M) such that the sur- inweights M ,1]upLet toby isotopy. A refinement [O] is that the Bi ’s ∂ incompressible Theorem [29, Theorem M be a Haken 3-manifoldin(compact, orientable, faces carried 3.1.5. with surfaces positive these B i ’s are exactly all the incompressible, can be chosen so that all the surfaces they carry, whether of positive weights or not, surfaces in Ma , two-sided up to isotopy. A refinement [O] is that the Bi ’s ∂ incompressible irreducible 3-manifold containing incompressible surface) with in incompressible are incompressible and ∂ incompressible. can be chosen so that all the surfaces they carry, whether of positive weights or not, boundary. arethese a finite number surfaces of properlyBembedded branched surfaces B1 , . . . ,track, Bk , Let B beThere one of branched i . Then ∂B = B ∩ ∂M is a train are incompressible and ∂ incompressible. ∂M with twoin key or such branched 1 manifold, that every two-sided in essential surface M properties: is isotopic to a surface carried by one of Let B be one of these branched surfaces Bi . Then ∂B = B ∩ ∂M is a train track, (1)theThere Bi ’s. is no smooth disk D ⊂ ∂M with D ∩ ∂B = ∂D . or branched 1 manifold, in ∂M with two key properties: (2) There is no disk D ⊂ ∂M , smooth except for one outward cusp point in ∂D , such Their [29] also proof Hatcher’s result that the. list of boundary slopes that Dpaper ∩ = ∂D . contains (1) There is no∂Bsmooth disk Da ⊂ ∂Mof with D∩ ∂B = ∂D is finite for a link using surfaces [54]. (2)The There no disk D ⊂ ∂M , smooth except one outward cusp point latteris condition is branched explicitly given in [FO]. Iffor condition (1) failed, then any in sur-∂D , such face carried that D ∩ by ∂B B=with ∂D .positive weights would have a boundary circle which was conTheorem 3.1.6. Corollary to Theorem For awould knot Kbound ⊂ S 3 , athere only finitelyof tractible in ∂M . By[54, incompressibility, this 1] circle diskare component Thethelatter condition explicitly givenofinB [FO]. condition thenasany sursurface, contrary is to the construction in [FO].IfCondition (2)(1) canfailed, be phrased 3 many slopes realized by boundary curves of essential surfaces in S K. saying thatby theBtrain track ∂B hasweights no monogons. train tracks are required face carried with positive wouldSometimes have a boundary circle which was con- to have no digons well, but we have to allow these here. tractible ∂M . Byasincompressibility, this circle would bound a disk of Wein study branched surfaces in the specific setting of Montesinos knots, which are component knots Let S be a surface by B with positive weights. No component of ∂S the surface, contrary tocarried the construction of B in [FO]. Condition (2) can be can phrased as obtained by connecting rational tangles in a cyclic pattern, see Figure 3.4. Let B be a 3be in ∂M , since there would be a smooth disk ⊂ ∂M are withrequired sayingcontractible that the train track ∂Botherwise has no monogons. Sometimes trainDtracks with foursomewhere points marked on itsthis boundary. A (2-)tangle is (2) a pair of properly embedded ∂Dball ⊂ ∂B , and inside disk condition (1) or would be violated. Thus to have no digons as well, but we have to allow these here. in each component torus Ti of ∂M which B meets, ∂S consists of a number of parallel Let S be a surface carried by B with positive weights. No component of ∂S can circles. 64 be contractible ∂M , since otherwise there be a smooth disk one of the tori of ∂M we let D ∂i B⊂=∂M with To simplifyinnotation in what follows, if Ti is would ∂D∂B ⊂∩ ∂BTi, ,and somewhere inside disk condition (1) or (2) would be violated. Thus and similarly ∂i S = ∂S ∩ Tthis i for any surface S carried by B . arcs into B each endpoint of the arcs is sent to a marked point on the boundary. A tangle consisting of two unlinked arcs is a trivial tangle. Definition 3.1.7. A tangle is a rational tangle if one can obtain a trivial tangle after applying a finite number of consecutive twists to neighboring endpoints of the tangle, see Figure 3.4a. To each rational tangle we associate a slope p/q. See [21] for Conway’s original treatment on rational tangles, where he proves that two rational tangles are isotopic if and only if their slopes are the same. p1 q1 (a) A rational tangle. p2 q2 pn−1 qn−1 pn qn (b) A Montesinos knot K(p1 /q1 , p1 /q2 , ..., pn /qn ). Figure 3.4: Schematics for a Montesinos knot. We denote a Montesinos knot by K(p1 /q1 , p2 /q2 , . . . , pn /qn ) which indicates that the link is obtained by connecting n rational tangles of slope pi /qi . A pretzel knot is a Montesinos knot of the form K(1/q1 , 1/q2 , . . . , 1/qn ). We will just write P (1/q1 , 1/q2 , . . . , 1/qn ) for pretzel knots from now on. The Hatcher-Oertel algorithm is obtained by patching together essential surfaces for each rational tangle, which we discuss in the next section. 3.2 Essential surfaces in 2-bridge knot complements Definition 3.2.1. A rational, or 2-bridge knot K(p/q), is a knot obtained by closing a single rational tangle of slope p/q with two trivial arcs. 65 .... ,n~_l) of n parallel sheets running the vertical porS.(n~ ....S.(n~ ,n~_l) consistsconsists of n parallel sheets running close toclose the to vertical porn d 1of 1..... bk], bifurcate which bifurcate n~ parallel copies tions oftions each ofb aeach n d ofb aE[-b .....E[-b bk], which into n~ into parallel copies of the of the b i n g square n - n i parallel outer plumbing i'h inneri'h pinner l u m b i npgl u m square and n - nandi parallel copies copies of the of ith the outerith plumbing F o r example, n = surfaces 1, the surfaces ( h i =1)0 are or 1) are square. square. F o r example, when nwhen = 1, the S l(n 1. . .S. l(n , n k 1- .1.). . (, nhki =- 10) or 2 k- ~ plumbings of the original just thejust 2 k- the ~ plumbings of the original k bands.k bands. the surfaces ,nk_different 0 for different 1.... ,bk]'s In orderIn toorder havetoallhave the all surfaces S,(n~ ....S,(n~ ,nk_.... 0 for Z'[b 1....Z'[b ,bk]'s in a copy singleofcopy S 3 -weKp/q, we choose fixed position forsay Kp/q, lying inlying a single S 3 - of Kp/q, choose a fixed aposition for Kp/q, the say the Incompressible surfaces in 2-bridge knot complements A. Hatcher 1 and W. Thurston 2 Department of Mathematics, Cornell University, Ithaka, NY 14853, USA Department of Mathematics, Princeton University, Princeton NJ 08544, USA To each rational number Kp/q shown in Fig. 1. QI p/q, with q odd, there is associated the 2-bridge knot bl (a) The “two bridges.” (b) 4-plat.Fig. 2 Fig. 1. The 2-bridge knot Kp/q Fig.(c) 2 A branched surface obtained by plumbing. In (a), the central grid consists of lines of slope +p/q, which one can magine as being drawn on a square "pillowcase". In (b) this "pillowcase" is punctured and flattened out onto a plane, making the two "bridges" more vident. The knot drawn is K3/5, which happens to be the figure knot. Figure 3.5: eight Figure from [56]. We assume q odd in order to get a knot rather than a two-component link.) The double cover of S 3 branched along Kp/q is the lens space Lq,p. With this observation, attributed in [16] to Seifert, the isotopy classification of 2-bridge knots follows easily from the classification [14]well-studied. of oriented lens spaces: K~/q Two-bridge knots are They have a 4-plat representation, see Figure 3.5b, =gp,/q, if and only if q'=q and p,_p+_l (modq). Basic references for 2-bridge Fig. 3 Fig. 3 knots are [2, 16, 17]. and they have been classified by Schubert [103] via the classification of lens spaces. In [56], We shall derive in this paper the isotopy classification of the incompressible urfaces, orientable or not, i n S3-Kp/q. As an application, we obtain some and Thurston presents an algorithm enumerate all boundary slopes of a 2-bridge nformation aboutHatcher the manifold resulting from Dehn surgery o n Kp/q:to Excludng the cases when Kp/q is a torus knot (Dehn surgery on torus knots was completely analyzed in [10]), every surgery K m yields an irreducible knot K(p/q) byDehn finding all on possible continued fraction expansions of p/q, each of which manifold, and all but finitely many Dehn surgeries yield non-Haken (i.e., not ufficiently large) corresponds non-Seifert-fibered manifolds. surface The caseinofthe the complement figure eight of K(p/q). to a branched I I Let p/q = r + [b1 , . . . , bk ] correspond to a continued fraction expansion of p/q. 1 p/q = r + b1 − . 1 b2 − · · · From each continued fraction expansion, we get a branched surface Σ[b1 , . . . , bk ] obtained by plumbing k bands, each having bi twists (right-handed if bi > 0 and left-handed if 66 bi > 0), using both the inner and outer plumbing squares. See Figure 3.5c. We denote by Sm (n1 , . . . , nk−1 ) the surface carried by Σ[b1 , . . . , bk ] consisting of m parallel sheets running along vertical portions of each band, which split into ni parallel copies of the ith inner plumbing square and n − ni parallel copies of the ith outer plumbing square. Theorem 3.2.2. [56, Theorem 1(b)] An essential surface with boundary in S 3 K(p/q) is isotopic to one of the surfaces Sm (n1 , . . . , nk−1 ) carried by Σ[b1 , . . . , bk ]. We describe part of the proof of the theorem, particularly the notion of edgepaths which will be relevant for the description of the Hatcher-Oertel algorithm. To prove Theorem 3.2.2, Hatcher and Oertel use the diagram in Figure 3.5a with the height function regarded as a natural projection S 3 → R, so that the knot lies in S 2 × [0, 1]. ˚2 The complement of the knot in each level sphere Sr2 = S 2 × {r} is a 4-punctured sphere S r for 0 < r < 1. Consider the quotient R2 /Γ, where Γ is the set of 180◦ rotations about the integer lattice. This covering of S 2 by R2 branches at the four punctures. We may regard (a) isotopy classes of smooth circles in Sr2 separating the four punctures into pairs, and (b) isotopy classes of smooth arcs in Sr2 joining one given puncture to any of the other three punctures as lines in R2 , by considering its preimage in the covering. Thus we can assign the slope of ˚2 . Now K(p/q) ∩ S 2 with the diagram the respective line in R2 to each circle and arc on S r r that we have chosen consists of two arcs of slope 1/0 for r = 1 and two arcs of slope p/q for r = 0. The height function on an essential surface S may be regarded as a Morse function with critical points at distinct r’s. Using the incompressibility and the ∂-incompressibility of S, 67 the intersection of S with each level sphere Sr2 consist of parallel copies of circles and arcs of the same slopes joining a pairs of punctures, or two sets of arcs, each of a different slope, joining different pairs of punctures. See Figure 3.6. n n n n m Figure 3.6: Intersection of S with Sr2 [56]. We represent a curve system of arcs of two different slopes a/b and c/d by k a m−k + m b m c , d so that there are 2k arcs of S ∩ Sr2 of slope a/b and 2(k − m) arcs of slope c/d. Definition 3.2.3. Let S be a connected and properly embedded surface in S 3 K. The number of intersections of S with the meridian circle of N (K) is called the number of sheets of S. Since S is assumed to be properly embedded, at level r = 0, S ∩ Sr2 consists of arcs of a single slope joining two punctures, and at level r = 1, S ∩ Sr2 are arcs of ∞ slope. Going from r = 0 to r = 1, the surface S must pass through a sequence of saddles by which the slopes of the arcs and circles changes to a different one. Suppose that a saddle changes arcs of slope a/b to slope c/d. Through a linear change of coordinate, these arcs may be brought to arcs of slope 0/1 and 1/0. Thus, ad − bc = 1. Consider the Poincare disc model of the hyperbolic plane with rational points marked on the boundary. We define a 1-simplex C on the hyperbolic plane by letting the vertices 68 be the rational points on the boundary. Additionally, there is an edge between fractions a/b and c/d if ad − bc = 1. See Figure 3.7. 2/1 5/3 3/2 4/3 1/1 3/4 2/3 3/5 1/2 5/2 2/5 3/1 1/3 4/1 1/4 1/0 0/1 -4/1 -1/4 -1/3 -3/1 -2/5 -5/2 -2/1 -5/3 -2/3 -3/2 -4/3 -1/1 -3/4 -1/2 -3/5 Figure 3.7: The 1-simplex on the Poincare disc model [56]. Definition 3.2.4. An edgepath of C is a path on the 1-simplex C. We represent the sequence of saddles of S by an edgepath, where an edge corresponds to a saddle passing between two slopes. For S to be essential, the edgepath needs to satisfy the following: (E1) no three successive vertices of the edgepath lie on two different edges of the same triangle of the diagram. (E2) pi /qi = pi+2 /qi+2 for each i. In other words, no backtracking is allowed. Given any continued fraction expansion p/q = r + [b1 , . . . , bk ], we have that the sequence 69 of partial sums pi /qi = r + [b1 , . . . , bi ], 1 ≤ r ≤ k satisfies pi qi+1 − qi pi+1 = ±1. Conversely, suppose that we have a sequence of fractions {pi /qi } satisfying the equation for pi /qi and pi+1 /qi+1 , the conditions (E1) and (E2), and such that pk /qk = p/q and p1 /q1 = {−1, 0, 1}. Then, we can construct a partial fraction decomposition of p/q. Thus an essential surface for K(p/q) corresponds to a continued fraction expansion of p/q. To find the branched surface carrying S, take the continued fraction expansion r + [b1 , . . . , bk ] of p/q given by the edgepath of S. Isotope the branched surface Σ[b1 , . . . , bk ] so that its boundary is the diagram 3.5a of K(p/q). It is clear that S is carried by Σ[b1 , . . . , bk ]. 3.3 Curve systems and boundary slopes for Montesinos knots Viewing S 3 as the join of two circles A and B, let the circle B be subdivided as an n-sided polygon. The join of A with the ith edge of B is then a ball Bi . We choose Bi so that each of them contains a tangle of slope pi /qi . These n balls Bi cover S 3 , meeting each other only in their boundary spheres. The union of the tangles in each ball with the trivial tangle gives K(p1 /q1 , p2 /q2 , . . . , pn /qn ). Within each ball Bi , we can still view the tangle as lying in S 2 × [0, 1], as before. The intersection of an essential surface S of K(p1 /q1 , p2 /q2 , . . . , pn /qn ) with a level sphere Sr2 will determine an edgepath for each fraction pi /qi satisfying the same criteria E(1), E(2). Since we can determine the essential surface for each tangle, we can reverse the process and 70 look for conditions that will allow them to fit together to make an essential surface for the Montesinos knot. The key difference between the case for 2-bridge knots and for Montesinos knots now is that at r = 1, the surface S determines a curve system on the 4-punctured S12 which does not necessarily consist solely of slope 1/0 arcs. With certain conditions, the curve system of each surface may be identified with each other to give a connected, essential surface. Definition 3.3.1. On a compact surface (with or without boundary), a curve system is a finite collection of disjoint, embedded curves which are either • circles not bounding disks in the surface and not isotopic to boundary components of the surface. • arcs with endpoints on the boundary components of the surface and not isotopic rel endpoints to arcs in the boundary. We will consider the curves a, b, and c on the 4-punctured sphere, where a is an arc joining two punctures, b is a circle containing the two punctures joined by a, and c is a circle containing one of the puctures contained by b and one of the other punctures. c a a c b c a b c Figure 3.8: The generators a, b and c and the corresponding set of disjoint curves on the 4-punctured sphere with a, b, c-coordinates (3, 1, 2). 71 ˚2 denotes the 4-punctured S 2 . We consider the subgroup of H1 (S ˚2 ) generRecall that S ated by a, b, and c, quotiented out by multiples of the coefficients of a, b, and c. This is a ˚2 . subset of the projective lamination space on S As discussed in the previous section, a curve system determined by an essential surface on ˚2 can be uniquely expressed in the form na a + nb b + nc c, where na are the coefficients of the S ˚2 ). We will write (na , nb , nc ) to indicate the curve system. We have the curve system in H1 (S planar three-simplex shown below where each point represents, in barycentric coordinate, a curve system (na , nb , nc ), where nc > 0. c 3/1 3/1 2/1 2/1 1/1 1/1 1/2 0/1 a 0/1 b Figure 3.9: The three-simplex of curve systems on a four-punctured sphere [55]. ˚2 in R2 Γ, where Γ is the integer Recall that by considering the preimage of a curve on S lattice in the plane, we may assign a slope to the curve. Choose coordinates on R2 so that a and b have slop 0/1 and c has slope 1/0. For a curve system represented by (na , nb , nc ), the slope is given by nc /(na + nb ), which we mark next to each vertex. There is also a mirror image of the simplex with c < 0 carrying curve systems of negative slopes. Joining the left-hand edges of both of those 2-simplices to the point representing the slope 1/0 arcs joining the other pairs of punctures gives a planar four-simplex. Now we see the familiar figure of Figure 3.7 in the form of a planar four-simplex. In 72 1/1 3/2 2/1 3/1 2/3 1/2 1/3 1/0 0/1 −3/2 −1/3 −1/2 −2/3 −2/1 −3/1 Figure 3.10: The four-simplex of curve systems with systems of slope 1/0 added [55]. addition to edges between slopes p/q and r/s where ps − rq = ±1, an edge between two points with the same slope p/q is allowed. An edgepath of Figure 3.7 is also an edgepath on this new diagram. For each ball Bi containing the tangle of slope pi /qi , the axis A is a slope 1/0 circle, and the circle which is the joint of two points of B ∩ ∂Bi with two antipodal points of A is a slope 0 circle. In the complement of K(p1 /q1 , p2 /q2 , . . . , pn /qn ), the right hemisphere of Bi as divided by the axis is identified with the left hemisphere of Bi+1 , forming a new curve system on the boundary sphere of the union. Suppose we have the curve system (na , nb , nc ) on ∂Bi and the curve system (na , nb , nc ) on ∂Bi+1 . In view of homology, the necessary condition to add them is that na = na , nb = nb . The new curve system will have the same coefficients for a and b while the coefficients for c are added. Now we slit the 1-simplex C open along the slope 1/0 edge, see Fig 3.11. The 1-simplex now forms an infinite strip D on the plane. Each point (na , nb , nc ) now has horizontal coordinate nb /(na + nb ) and vertical coordinate nc /(na + nb ). An edgepath from the point 73 1/1 2/1 3/4 2/3 3/5 1/2 1/1 ⇒ 1/0 1/3 1/6 1/5 0/1 −1/1 1/1 0/1 0/1 Figure 3.11: On the left: the infinite strip containing the 1-dimensional complex D. On the right: the portion of the infinite strip from height 0/1 to 1/1 [55]. of slope p/q to the point of slope r/s corresponding to a saddle within a ball Bi from curve systems of slope p/q to the system of slope r/s is now represented as p/q − r/s . We will decorate the point corresponding to the curve system (0, q, p) with slope p/q with the symbol ◦ to distinguish it from the curve system (1, q − 1, p) with the same slope. For an edge p/q − r/s , let k/m p/q + (m − k)/m r/s denote the curve system consisting of k parallel copies of the pair of arcs of slope p/q joining the four punctures together with m − k parallel copies of the pair of arcs of slope r/s. This is a point on p/q − r/s . In terms of a, b, and c, the curve system has coordinates (m, k(q − 1) + (m − k)(s − 1), kp + (m − k)r). As before, we isotope an essential surface so that its height function is a Morse function. 74 The sequence of saddles in each ball Bi correspond to an edgepath on the complex D. Therefore, we may construct a candidate surface by considering the choices of edgepath for the tangle in each ball, and ask whether the surfaces associated to the edgepath glue together to form a connected, properly embedded, and essential surface. The endpoint of an edgepath will correspond to a curve system on the 4-punctured sphere which we glue with the curve system on the other 4-punctured sphere. In order for the pieces of these surfaces to form a connected surface after identifying Bi ’s, we need to have na = na , nb = nb as before. This means that the horizontal coordinates of the endpoints of each edgepath in D need to agree, and the vertical coordinates need to sum to zero. We can now describe the Hatcher-Oertel algorithm for enumerating all candidate surfaces for boundary slopes. 3.4 The Hatcher-Oertel algorithm For each fraction pi /qi , an edgepath γi is a piecewise linear path in the 1-skeleton of D, starting and ending at points which may not be vertices of D. It satisfies the following conditions: (E1) The starting point of γi lies on the edge pi /qi − pi /qi◦ . If this starting point is not a vertex, then γi is constant. (E2) γi never stops and retraces itself or go along two sides of the same triangle in D in succession. This is the same local condition coming from incompressibility and ∂incompressibility as in Section 3.2 for 2-bridge knots. 75 (E3) γi proceeds monotonically from right to left where motion along vertical edges are permitted. We are looking for an edgepath system {γ1 , . . . , γn } satisfying the additional properties from the conditions on identifying curve systems. (E4) The endpoints of all the γi ’s are rational points of D which have the same horizontal coordinates, and whose vertical coordinates add up to zero. We associate to an edgepath system {γ1 , . . . , γn } a candidate surface as follows. Viewing each tangle pi /qi as lying in S 2 × [0, 1] in Bi as before. There are two cases for γi . • If γi is constant: We associate the surface Si meeting each level Sr2 in the curve system with m arcs of slope p/q coming into each puncture. The circles in the intersection Si ∩ S02 are capped off with disks. • If γi is not constant: The surface is associated just as before in the case for 2-bridge knot. If an edgepath ends at the point with slope 1/0, then all the other edgepaths in the system also have to end at the same point. To enumerate all the edgepath systems given a Montesinos knot K(p1 /qi , . . . , pn /qn ), one finds all possible edgepaths. 1. Each γi either correspond to a continued fraction expansion of pi /qi or has a constant edgepath p/q − p/q ◦ added at the beginning. 2. Given a choice of edgepath system {γ1 , . . . , γn }, solve for endpoints satisfying condition (E4). 76 Each of the edgepath systems with the appropriate endpoints then give a candidate egdepath system. Depending on the fractions pi /qi , some edgepath systems will not have endpoints satisfying (E4), and they will be ruled out. Hatcher and Oertel shows that the algorithm gives a complete list of candidate surfaces for boundary slopes. Theorem 3.4.1. [55, Proposition 1.1] Every essential surface in S 3 − K(p1 /q1 , . . . , pn /qn ) having nonempty boundary of finite slope is isotopic to one of the candidate surfaces. Just as in the case for 2-bridge knots, we obtain a branched surface in the complement of the Montesinos knot by gluing branch surfaces in the complement of each tangle. Every essential surface is carried by a branch surface constructed in this way. Thus the surfaces of the edgepaths systems are similar to that of the case for two-bridge knots, except that a constant edge is allowed at the beginning of an edgepath. The endpoint of an edgepath in a candidate edgepath system is also not required to end at a curve system of slope 1/0. To check that a given candidate surface is essential, Hatcher and Oertel used a technical idea of the r-values of the edgepath system, which we will not describe in more detail in this dissertation. The idea is to examine the intersection of a compressing or ∂-compressing disk with the boundary sphere of each ball Bi , which will determine an r-value for each edgepath γi . If a candidate surface does not have a list of r-values from its edgepath system that would result from the existence of a compressing disk, then it is incompressible. We state the criterion in terms of quantities thare are easily computed given an edgepath system. Definition 3.4.2. The r-value for an edge p/q - r/s where p/q = r/s is s−q. If p/q = r/s 77 or the path is vertical, then the r-value is 0. Then the r-value for an edgepath γi is just the r-value of the final edge of γ. Theorem 3.4.3. [55, Corollary 2.4] A candidate surface is incompressible unless the cycle of r-values of {γi } is one of the folloing types: • (0, r2 , . . . , rn ), or • (1, . . . , 1, rn ). Dunfield has implemented the algorithm completely in a program [27] that will determine the list of boundary slopes given any Montesinos’ knot. 3.4.1 Computing the boundary slope from an edgepath system Given an edgepath system {γi } corresponding to an essential candidate surface, we describe how to compute its boundary slope. Within a ball Bi , all surfaces look alike near S02 , since it must have slope p/q at the bottom. The number of times the boundary of the surface winds around the longitude is given by m, the number of sheets of the surface. Each time the surface passes through a slope-changing saddle, the boundary twists around the meridian once. Thus the total number of twists for a candidate surface S in arcs away from the level r = 0 is τ (s) = 2(s− − s+ ), where s− is the number of slope-decreasing saddles and s+ is the number of slope-increasing saddles. In terms of egdepaths, we have τ (S) = 2(e− − e+ ), 78 where e− is the number of edges of γi that decreases slope, and e+ is the number of edges that increases slope. If γi ends at a point on the segment k m p q + m−k m r , s then the final edge is counted as a fraction k/m. We add back the twists happening around the knot at level r = 0 by subtracting the twist number of a Seifert surface S0 obtained from the algorithm, since the Seifert surface will always have zero boundary slope. The boundary slope of a candidate surface S is τ (S) − τ (S0 ). 3.4.2 The 3-strand pretzel knots P (−1/r, 1/s, 1/t). For Theorem 1.1.2, we shall restrict the Hatcher-Oertel algorithm to 3-string pretzel knots P (−1/r, ±1/s, ±1/t), where r is even and s, t are odd. We restrict the algorithm to this case due to the relative ease in computing the corresponding Jones slope, which we will describe in the next section. For each fraction of the form 1/p, there are two choices of edgepaths that satisfy conditions (E1)-(E3). They correspond to the two continued fraction expansions of 1/p: 1 = 0 + [p] gives edgepath p 1 p − 0, 79 and 1 = ±1 + [±2, ±2, . . . , ±2] gives edgepath p p − 1 times ±1 p − ±1 p−1 − · · · − ±1 . where it is a plus or minus sign for the slope of each vertex for the second type of continued fraction expansion depending on whether p is positive or negative, respectively. These edgepaths are shown on the diagram D. 1/1 3/4◦ 1/2◦ 1/4◦ 1/8◦ 0/1 Figure 3.12: A choice of edgepath for each of the three fractions 1/2, 1/8, and 3/4 are shown in red, blue, and green. Thus for the 3-strings pretzel knots there are 23 choices of edgepath systems, which we may augment with constant edgepaths at the beginning. Finding a candidate edgepath system amounts to solving the equations for the horizontal and vertical coordinates imposed by condition (E4). We consider the following edgepath system for P (−1/r, 1/s, 1/t) when r > s, t > 0. • For 1/r: −1 r − −1 r−1 80 − · · · − −1 . • For 1/s: 1 s − 0 . 1 t − 0 . • For 1/t: 0 − 1/t Condition (E3) requires that we set the horizontal coordinates of points on the path of form −1/(q + 1) − −1/q , and 1/s − 0 , and 1/t − 0 equal to each other: That is, we have the following equation. m(q − 1) + k k (s − 1) k (t − 1) = = . mq + k m + k (s − 1) m + k (t − 1) Recall that for the curve system represented by each point, the number k represents the number of arcs coming into each puncture with one slope and m − k represents the number of arcs coming into each punctre of a different slope. The number of sheets m will be the same. Dividing the top and the bottom by m, we get k /m(s − 1) k /m(t − 1) (q − 1) + k/m = = . q + k/m 1 + k /m(s − 1) 1 + k /m(t − 1) (3.1) We simplify the notation by setting A = k/m, B = k /m, and C = k /m. q−1+A B(s − 1) C(t − 1) = = . q+A 1 + B(s − 1) 1 + C(t − 1) 1 1 1 = = . q+A 1 + B(s − 1) 1 + C(t − 1) q + A = 1 + B(s − 1) = 1 + C(t − 1). 81 (3.2) The sum of the vertical coordinates are equal to zero, which means we have k k −m + + = 0. mq + k m + k (s − 1) m + k (t − 1) (3.3) Since all the denominators are the same, we also get −m + k + k = 0, and −1 + B + C = 0. (3.4) By equation (3.2), B= C(t − 1) . s−1 (3.5) We can then solve for c in terms of given quantities s, t using equation (3.4), obtaining C= s−1 . t+s−2 (3.6) 82 Now the boundary slope is given by τ (S) − τ (S0 ), where S0 is a candidate surface that is a Seifert surface. We can describe the edgepath system for S0 . • For −1/r: −1/r − 0 . • For 1/s: 1/s − 1/(s − 1) − · · · − 1/2 − 1 . • For 1/t: 1 − 1/2 − · · · 1/(t − 1) − 1/t . corresponding to the edgepath system with the number of sheets m = 1, and each γi turning across an even number of triangles at each vertex in order for S0 to be orientable. For a general algorithm for determining a Seifert surface which is a candidate surface, see the discussion in [55], where the formula for the boundary slope is given. So τ (S0 ) = −2(−1 + s + t). To compute τ (S), note that the edges are all decreasing. We add up A, B, C and the number of paths for τ (S). We have, τ (S) = 2(r − q − A + 1 − B + 1 − C) = 2(r − q − A + 1). Given that q + A = 1 + C(t − 1), we get τ (S) = 2(r − (1 + C(t − 1)) + 1), so then τ (S) − τ (S0 ) = 2 r + t + (−1 + s)2 −2 + s + t . (3.7) The cycle of r-values is (1, s−1, t−1). Therefore by Theorem 3.4.3, the surface is essential 83 as long as s > 2. 3.5 Computation of the Jones Slope We describe the computation for the Jones slope done by Roland van der Veen, which is relatively simple to determine for the class of 3-string pretzel knots P (−1/r, 1/s, 1/t) where r > s, t and r, s, t > 5. We assume that r is even and s, t are odd as in the premise of Theorem 1.1.2. Recall from Section 1.2 that we may obtain the colored Jones polynomial by evaluating the Kauffman bracket on a link diagram decorated with the Jones-Wenzl idempotent. We list here a few important formulas for the graphical calculus with the idempotent. See [78] and [82] for more details and proofs. First we represent a special skein element, shown in Figure 3.13, by a trivalent graph on the right. a b a b z x y c c Figure 3.13: A trivalent graph representation [78]. The next few figures give formulas for replacing certain skein elements with a sum and the 6j-symbol. For k an integer, recall that we have k = (−1)k+1 (A2(k+2) − A−2(k+2)) . A2 − A−2 This is the skein element in S(R2 ) obtained by closing up fk with k parallel strands. We 84 also define x+y+z ! θ(α, β, γ) = where k! = k k−1 · · · 1 y+z−1 ! x−1 ! y−1 ! z+x−1 ! z−1 ! x+y−1 ! , and is interpreted as 1 if k is 0. We define the 6j-symbol, see [78] and [82] for references. Let a+b+e a+c+f b+d+f c+d+e , , , 2 2 2 2 zmin := max a+d+b+c a+d+e+f b+c+e+f , , 2 2 2 zmax := min     a b c     d e f   zmin ≤z≤zmax ! z− a+b+e 2 z+1 ! a+c+f ! z− 2 b+d+f ! z− 2 1 ! b+c+e+f −z 2 ! a+d+e+f −z 2 b+c+a+d −z ! 2 = c c:(a,b,c) θ(a,b,c) admissible c b Figure 3.14: The fusion formula [78]. a b = (−1) a+b−c a+b−c+ a2 +b2 −c2 2 A 2 Figure 3.15: The untwisting formula [78]. 85 ! z− c+d+e 2 . a a b c . := (−1)z · , and a b · b c b a b c d a = i c b a b ⇢ i P c d j a b i = i c d j d c i a i d a d Figure 3.16: The 6j-equation [78]. Figure 3.16: The 6j-equation [78]. c a d a ad ✓(a,b,c) 4a = b Figure 3.17: Replacing a region bounded by two edges. is the Kronecker delta function Figure 3.17: Replacing a region bounded by two edges. δad is the ad Kronecker delta function [78]. [78]. Since fn2 = fn , we may insert as many projectors for a link component. We may arrange Since fn2 = fn , we may insert as many projectors for a link component. We may arrange them so that for twist region, there is a Jones-Wenzl idempotent on each outgoing strands. them so that for twist region, there is a Jones-Wenzl idempotent on each outgoing strands. The e↵ect of replacing a positive crossing with a negative crossing is to replace A by A 1 . The effect of replacing a positive crossing with a negative crossing is to replace A by A−1 . Let d(↵, n) = 2n ↵ + n2 ↵2 and d( ) and d( ) be defined similarly. Using the fusion and Let d(α, n) = 2n − α + n2 − α2 and d(β) and d(γ) be defined similarly. Using the fusion and the untwisting formulas, the n-th colored Jones polynomial before substituting q 1/2 = A 1/2 the untwisting formulas, the n-th colored Jones polynomial before substituting q 1/2 = A−1/2 is then a sum: is then a sum: ✓ n2 +2n (2 +2n 1)n A−ω(D) n n (−1) A ◆ !(D) X 4 4 4↵ · β γ · n, ) ✓(n, n, ↵) ✓(n, n, ) ✓(n, θ(n, n, α) θ(n, n, β) θ(n, n, γ) ↵, , admissible α α,β,γ admissible * ⇣ ⌘ ⇣ ⌘ ⇣ ⌘ ↵ +s n r n +t n 2 A rd(↵,n)+sd( ,n)+td( ,n) · ( α2 1)+s n−2β2 +t n− 2γ2 −rd(α,n)+sd(β,n)+td(γ,n) r n− · (−1) α A 86 86 β + ↵ γ . . (3.8) (3.8) A B F E D = A B E D C F A 1 · θ(A,F,C) C F C Figure 3.18: Collapsing a triangular region to a vertex [82]. The sum is taken over positive integers α, β, and γ which are admissible. This means that each of them satisfy the admissibility conditions with respect to n. We remark that this is a restriction of the more general admissibility condition defined by Lickorish in [78] shown below α, β, and γ are even, 0 ≤ α, β, and γ ≤ 2n, and |α − β| ≤ γ ≤ α + β. (3.9) 87 For evaluating the colored Jones polynomial of general links, we may use the 6j-equation to reduce the number of edges around a complementary region of the trivalent graph. If a region is bounded by just two edges, we use the identity of Figure 3.17. There is a simple tetrahedral shape of the resulting trivalent graph for the case that we consider for the 3-string pretzel knot. We use the identity of Figure 3.18 to change the previous sum to 2 (−1)n An +2n −ω(D) α,β,γ admissible β γ r n− α 2 +s n− 2 +t n− 2 · (−1) · β α γ θ(n, n, α) θ(n, n, β) θ(n, n, γ) · A−rd(α,n)+sd(β,n)+td(γ,n) ·  2   α n n   α 1 2   n β γ   θ(α, β, γ) γ β (3.10) . n before Finally, using the symmetry relation of the 6j-symbol [18, 82], we have that JK subsituting q is 2 (−1)n An +2n ω(D) α α,β,γ admissible β γ θ(n, n, α) θ(n, n, β) θ(n, n, γ) 2 2 2 2 · (−1)6n−(α+β+γ)/2 A6n+3n −(α +β +γ )/2 ·  2   n n n   1 .   α β γ   θ(α, β, γ) n . We maximizes MD Let M D(α, β, γ) be the degree of a single term in the sum for JK over the 3-dimensional polytope defined by admissible parameters α, β, and γ. The solutions will be integer lattice points of the polytope, see Figure 3.19. 88 1.0 0.5 0.0 1.0 0.0 0.0 0.5 1.0 Figure 3.19: The polytope cut out by equations α + β = γ, α + γ = β, β + γ = α. It has been rescaled by dividing by n. In the event that there are more than one set of parameters which maximizes M D, cancellations are possible. However, the restrictions placed on r, s, and t from Theorem 1.1.2 guarantees that there is a single maximum degree term. The formula for the 6j-symbol is complicated, however, analyzing M D with certain restrictions on α, β, and γ enables us to write down explicit formulas for the restrictions. We rescale α, β, and γ by replacing them with nα, nβ, and nγ.      (−4 + 2r)α2 − 2sβ 2 − 2tγ 2 + 4αβ + 4αγ − r + s + t for α ≥ β, γ.      M D(α, β, γ) = (−4 − 2s)β 2 + 2rα2 − 2tγ 2 + 4αβ + 4βγ − 4αγ − r + s + t for β ≥ α, γ.         (−4 − 2t)γ 2 − 2sβ 2 − 2tα2 + 4γα + 4βγ − 4αβ − r + s + t for γ ≥ α, β. We call the restrictions of M to these three domains M α, M β and M γ. We see that 89 ∂M α ∂α >0 ∂M α ∂β <0 ∂M β ∂α >0 ∂M β ∂β <0 ∂M γ ∂α >0 ∂M γ ∂β <0 ∂M α ∂γ ∂M β ∂c < 0, < 0, and ∂M γ ∂γ < 0. Therefore, the maximum must lie in the triangle α = β + γ, β + γ ≤ 1. Thus we need only to find the integer lattice maximum of M α(b + c, b, c) over the triangle. The determinant of the Hessian of M α is H(b, c) = st − r(−2 + s + t) − 1. If H < 0, then the maximum is attained at one of the edges of the triangle, so the problem reduces to maximizing function over the edges of the triangle. The critical point on −1+t and the slope is the diagonal is b0 = −2+s+t 2 − 2s + s2 − 2t + t2 + r(−2 + s + t) + W (r, s, t), −2 + s + t where W (r, s, t) is the writhe of the pretzel knot P (−1/r, 1/s, 1/t). 3.6 Proof of Theorem 1.1.2. Proof. From the previous section, the Jones slope is 2 − 2s + s2 − 2t + t2 + r(−2 + s + t) + r + s + t, −2 + s + t after ignoring all terms of order less than n2 and multiplying by 4. The slope of an essential surface in the complement of P (−1/r, 1/s, 1/t) is equal to (−1 + s)2 2 r+t+ −2 + s + t 90 , as we have seen in Section 3.4. These are the same, and we are done. 91 Chapter 4 Crosscap number of alternating links and the Jones polynomial Clark [20] showed that the crosscap number of any knot K, is bounded above in terms of the orientable genus of K, and Murakami and Yasuhara [89] showed that it is bounded above in terms of the crossing number of K. For an alternating link K, C(K) > 1, unless K is a (2, p) torus link. A lower bound for the 4-dimensional crosscap number of all knots, and thus for the 3-dimensional crosscap number, was given by Batson [13]. For alternating knots, however, this bound is non-positive. We restate Theorem 1.1.3, which is the main result of this chapter. Let JK (t) = αK tr + βK tr−1 + . . . + βK ts+1 + αK ts denote the Jones polynomial of K, so that r and s denote the highest and lowest power in t. Set TK := |βK | + βK , where βK and βK denote the second and penultimate coefficients of JK (t), respectively. Also let sK = r − s, denote the degree span of JK (t). Theorem 1.1.3. Let K be a non-split, prime alternating link with k-components and with 92 crosscap number C(K). Suppose that K is not a (2, p) torus link. We have TK 3 + 2 − k ≤ C(K) ≤ TK + 2 − k, where TK is as above, and · is the ceiling function that rounds up to the nearest integer. Furthermore, both bounds are sharp. Combining Theorem 1.1.3 with the results of [89] and [72], we have the following. Theorem 4.0.1. Let K be an alternating, non-torus knot with crosscap number C(K) and let TK be as above. We have TK 3 + 1 ≤ C(K) ≤ min TK + 1, sK 2 where sK denotes the degree span of JK (t), and · is the floor function that rounds up to the nearest integer. Furthermore, both bounds are sharp. By a result of Menasco [83] a link with a connected, prime alternating diagram that is not the standard diagram of a (2, p) torus link is non-split, prime and non-torus link. Hence the hypotheses of Theorem 1.1.3 are easily checked from alternating diagrams. Crowell [22] and Murasugi [90] have independently shown that the orientable genus of an alternating knot is equal to half the degree span of the Alexander polynomial of the knot. Theorem 4.0.1 can be thought of as the non-orientable analogue of this classical result. The oriented link genus has been well studied, and a general algorithm for calculation is known [5, 51]. For knots with low crossing number, effective computations can also be made from genus bounds coming from invariants such as the Alexander polynomial and the Heegaard Floer homology [19]. Crosscap numbers, however, are harder to compute. 93 Although the crosscap numbers of several special families of knots are known, see [61,64,107], no effective general method of calculation is known. Some progress in this direction was made by Burton [17] using normal surface theory and integer programming. Despite that, for the majority of prime knots up to twelve crossings the crosscap numbers are listed as unknown in KnotInfo [19]. For alternating links, a method to compute crosscap numbers was given by Adams and Kindred [1]. They showed that to compute the crosscap number of an alternating link it is enough to find the minimal crosscap number realized by state surfaces corresponding to alternating link diagrams. Note that for a link with n crossings, there correspond 2n state surfaces that, a priori one has to search and select one with minimal crosscap number. The Adams-Kindred algorithm of [1] cuts down significantly this number, but the number of surfaces that one has to deal with still grows as the number of the “non-bigon” regions of the alternating link does. The advantage of Theorem 1.1.3 is that it provides estimates that are easy to calculate from any alternating diagram and, as mentioned above, these estimates often compute the exact crosscap number. Indeed, the quantity TK is particularly easy to calculate from any alternating knot diagram as each of βK , βK can be calculated from the checkerboard graphs of the diagram [24]. Our lower bound improves the lower bound given in KnotInfo [19] for 1472 prime alternating knots for which the crosscap number is listed as unknown and for 262 of these knots our bounds determine the exact value of the crosscap number. See Section 4.4 for some examples of these knots. In Section 4.1 we define augmented links and state definitions and results from [37, 75] that are necessary to this chapter. In Section 4.2, first we define state surfaces and we recall the results of [1] that we use. Then we prove Theorems 1.1.3 and 4.0.1 in Section 4.3. In 94 Section 4.4, we calculate the crosscap numbers of infinite families of alternating knots for which the lower bound is sharp as well as for several knots up to 12 crossings. Finally, in Section 4.5, we discuss generalizations of our results outside the class of alternating links. The overall technique used to prove Theorem 1.1.3 goes beyond the class of alternating links and allow us to generalize Theorem 1.1.3 for large classes of non-alternating links. We also state some questions that arise from this work. 4.1 Augmented links and estimates with normal surfaces Consider a link diagram D(K) as a 4–valent graph in the plane, with over–under crossing information associated to each vertex. A bigon region is a region of the graph bounded by only two edges. A twist region of a diagram consists of maximal collections of bigon regions arranged end to end. We will assume that the crossings in each twist region occur in an alternating fashion. A single crossing adjacent to no bigons is also a twist region. A or B A B Figure 4.1: Twist reduced: A or B must be a string of bigons [37]. Definition 4.1.1. A link diagram D(K) is prime if any simple closed curve which meets 95 two edges of the diagram transversely bounds a region of the diagram with no crossings. The diagram D(K) is called twist reduced, if any simple closed curve that meets the diagram transversely in four edges, with two points of intersection adjacent to one crossing and the other two adjacent to another crossing, bounds a (possibly empty) collection of bigons arranged end to end between the crossings. See Figure 4.1, borrowed from [32]. 4.1.1 An angled polyhedral decomposition. For a link K ⊂ S 3 , let E(K) denote the exterior of K; that is E(K) = S 3 η(K). Let D(K) be a prime, twist–reduced diagram of a link K. For every twist region of D(K), we add an extra link component, called a crossing circle, that wraps around the two strands of the twist region. A choice of crossing links for each twist region of D(K) gives a new link J, called an augmented link. The link L obtained by removing all full twists from the twist regions of J is called a fully augmented link. The exteriors E(J), E(L) are homeomorphic 3-manifolds and E(K) can be expressed as a Dehn filling of E(L) ∼ = E(J). See Figure 4.2, borrowed from [32]. Figure 4.2: From left to right: A link K, an augmented link J and a fully augmented link L [32]. For twist regions consisting of a single crossing the addition of a crossing circle can be done in two ways. See Figure 4.3. The geometry of augmented links is well understood. Below, we will summarize some 96 Figure 2. A link K, an augmented link J and a fully augmented link L. Figure 3. Two ways of augmenting a twist region with a single crossing. Figure 4.3: Two ways of augmenting a twist region with a single crossing. the polyhedral decomposition and combinatorial length of simple closed curves that lie resultsto and properties we need; for more details see Purcell’s expository article [98] and the boundary @E(L). The general setting was discussed by Lackenby in [28] and applied by Futer and Purcell [17] to study the geometry of augmented highly twisted links. references therein. We will consider the non-crossing circle components of L flat on the proStart with a prime, twist reduced diagram of a link K and let L be a fully augmented link obtained from it. The manifold E(L) can be subdivided into two identical, convex jection plane and theP1crossing bound crossing disks vertical to the projection plane. ideal polyhedra and P2 circles in the hyperbolic 3-space. The ideal vertices of each polyhedron P 2 {P1 , P2 } are 4-valent and they correspond to crossing circles or to arcs of As explained by Agol and Thurston thecrossing appendix of [76], E(L) has an angled KrD, where D is theD.union of allinthe disks. After truncating all thepolyhedral ideal vertices of P we have: decomposition with nice geometric and combinatorial properties. We are particularly inter- • Dihedral angles at each edge of P are ⇡/2. In particular E(L) is hyperbolic. • There are two types of edges of @P : these that are created by truncation called ested in the combinatorial decomposition: define combinatorial boundary edges structure (edges of Pof\this @E(L)) and the ones one that can come from the edges existing before truncation, called interior edges. The interior edges come from intersecarea of surfaces that are in normal with plane. respect to the polyhedral decompotionsinofE(L) the crossing discs with theform projection • There are two types of faces on @P : these that are created by truncation called sition and combinatorial closedand curves thatthat lie come the boundary boundary faceslength (faces of of simple P \ @E(L)), the ones from faces∂E(L). existing The before truncation, called interior faces. • Each boundary facebyisLackenby a rectangleinthat four interior edgesand at the vertices general setting was discussed [75]meets and applied by Futer Purcell [37] to of the rectangle. The boundary faces of P1 , P2 subdivide @E(J) into rectangles. • The interior faces of P highly can be twisted colored with study the geometry of augmented links.two colors (shaded and white) in a checkerboard fashion so that at each rectangular boundary face of P opposite side interior twist faces have the same color. shaded faces to crossing Start with a prime, reduced diagram of The a link K and letcorrespond L be a fully augmented disks while the white faces correspond to regions of the projection plane. See Figure andmanifold Figure 12. link obtained[29, from it. 15] The E(L) can be subdivided into two identical, convex The complement of all faces (interior and boundary) on the projection plane is a graph ⇢ @P with at least three and edges edges of P .of each polyhedron ideal polyhedra P1vertices and P2ofinvalence the hyperbolic 3-space. The the ideal vertices A polyhedral P with the above properties is called a rectangular-cusped polyhedron. P ∈ {P1 , P2 } are 4-valent and they correspond to crossing circles or to arcs of K D, where D is the union of all the crossing disks. After truncating all the ideal vertices of P we have: • Dihedral angles at each edge of P are π/2. In particular E(L) is hyperbolic. • There are two types of edges of ∂P : these that are created by truncation called boundary edges (edges of P ∩ ∂E(L)) and the ones that come from edges existing before truncation, called interior edges. The interior edges come from intersections of the 97 crossing discs with the projection plane. • There are two types of faces on ∂P : these that are created by truncation called boundary faces (faces of P ∩ ∂E(L)), and the ones that come from faces existing before truncation, called interior faces. • Each boundary face is a rectangle that meets four interior edges at the vertices of the rectangle. The boundary faces of P1 , P2 subdivide ∂E(J) into rectangles. • The interior faces of P can be colored with two colors (shaded and white) in a checkerboard fashion so that at each rectangular boundary face of P opposite side interior faces have the same color. The shaded faces correspond to crossing disks while the white faces correspond to regions of the projection plane. See [76, Figure 15] and Figure 4.12. The complement of all faces (interior and boundary) on the projection plane is a graph Γ ⊂ ∂P with vertices of valence at least three and edges the edges of P . A polyhedral P with the above properties is called a rectangular-cusped polyhedron. 4.1.2 Normal surfaces and combinatorial area. We are interested in surfaces with boundary that are properly embedded in E(L) and are in normal form with respect to the above polyhedral decomposition. We recall the following definition. Definition 4.1.2. A properly embedded surface (F, ∂F ) ⊂ (E(L), ∂E(L)) is said to be in normal form with respect to the polyhedra decomposition, if for any P ∈ {P1 , P2 } we have the following: 98 1. F ∩ P consists of properly embedded disks (D, ∂D) ⊂ (P, ∂P ). 2. F ∩ ∂P is a collection of simple closed curves none of which lies entirely in a single face 6 E. KALFAGIANNI AND C. LEE of P . 2.2. Normal surfaces and combinatorial area. In this paper we are interested in surfaces with boundary that are properly embedded in E(L) and are in normal form respect to the of above decomposition. We recall the following 3. with F intersects faces P inpolyhedral a collection of properly embedded arcs none ofdefinition. which passes Definition 2.2. A properly embedded surface (F, @F ) ⇢ (E(L), @E(L)) is said to be in throughform vertices Γ. Furthermore, none decomposition, of these arcs runs from normal withofrespect to the polyhedra if for anyan P edge 2 {P1of , PF2 }to weitself have the following: or from edge to an adjacent boundary (1) Fan \ Pinterior consists of properly embedded disks (D,edge. @D) ⇢ (P, @P ). (2) F \ @P is a collection of simple closed curves none of which lies entirely in a single face of P . 4. A component of F ∩ P can intersect each boundary face in at most one arc. (3) F intersects faces of P in a collection of properly embedded arcs none of which passes through vertices of . Furthermore, none of these arcs runs from an edge of F to itself or from an interior edge to an adjacent boundary edge. The components of F ∩ Pofare An example normal disks in a (4) A component F \called P cannormal intersectdisks. each boundary face of in three at most one arc. The components \ P areincalled normal disks. An example of three in athere truncated polyhedronofisFshown Figure 4.4, borrowed from [37]. Note normal that P disks shown truncated polyhedron is shown in Figure 4, borrowed from [17]. Note that P shown there a rectangular-cusped polyhedron as not boundaryfaces facesare are rectangles. rectangles. is notisa not rectangular-cusped polyhedron as not allall thethe boundary Figure 4.4: Normal disks in aintruncated polyhedron Figure 4. Normal disks a truncated polyhedron.[37]. We recall the definition of the combinatorial area of a surface in normal form. Let (D, (P, @P ) be a normal in a polyhedron 2 {P1 , P2 }. WeDefinition recall the 2.3. definition of@D) the ⇢combinatorial area ofdisk a surface in normalP form. Suppose that D crosses m interior edges of P . The combinatorial area of D, denoted by a(D), is defined by m⇡ Definition 4.1.3. Let (D, ∂D) ⊂ (P, a\ normal in a polyhedron P ∈ {P1 , P2 }. a(D) = ∂P )+be ⇡|D @E(L)|disk2⇡, 2 where |D D \ @E(L)| the number of @D running on boundary faces of P . by Suppose that crosses denotes m interior edges ofof Parcs . The combinatorial area of D, denoted For an embedded surface (F, @F ) ⇢ (E(L), @E(L)), in normal form, the combinatorial area a(F ) is defined by summing over all the normal disks of F in the polyhedra P1 , P2 . a(D), is defined by Next we recall the following form the combinatorial version of the Gauss-Bonnet themπ orem due to Lackenby [28] a(D) = + π|D ∩ ∂E(L)| − 2π, 2 Proposition 2.4. [28, Proposition 4.3] Let (F, @F ) ⇢ (E(L), @E(L))) be a surface in normal form of Euler characteristic (F ). We have a(F ) =99 2⇡ (F ). where |D ∩ ∂E(L)| denotes the number of arcs of ∂D running on boundary faces of P . For an embedded surface (F, ∂F ) ⊂ (E(L), ∂E(L)), in normal form, the combinatorial area a(F ) is defined by summing over all the normal disks of F in the polyhedra P1 , P2 . Next we recall the following form the combinatorial version of the Gauss-Bonnet theorem due to Lackenby [75] Proposition 4.1.4. [75, Proposition 4.3] Let (F, ∂F ) ⊂ (E(L), ∂E(L))) be a surface in normal form of Euler characteristic χ(F ). We have a(F ) = −2πχ(F ). To continue, with P as above, consider a normal disk (D, ∂D) ⊂ (P, ∂P ) such that ∂D intersects at least one boundary face of P . Given an arc of γ ⊂ ∂D on a boundary face of P , define the combinatorial length of γ with respect to D by l(γ, D) = a(D) . |D ∩ ∂E(L)| For a simple closed curve γ ⊂ ∂(E(L), that is a boundary component of a surface (F, ∂F ) ⊂ (E(L), ∂E(L)), let H ⊂ F be the union of normal disks in F whose intersections with the boundary faces of P1 , P2 give γ. Thus γ is the union of arcs each properly embedded in a boundary face of the polyhedra. Now define the combinatorial length lc (γ) := l(γ, H) = i l(γi , D), where the sum is taken over all normal disks in H and all normal arcs on boundary faces. 100 The quantity lc (γ), defined above, depends on F . To obtain a well defined notion of combinatorial length one needs to consider the infimum over all normal surfaces F and collections H with ∂F = γ. In fact, one can define the combinatorial length of any simple closed curve on ∂E(L) [37, 75]. This is the definition of lc used by the authors in [37]. We will not repeat these definitions here as we don’t need them. However, the combinatorial length estimates of simple closed curves obtained in [37] also hold for l(γ, H). We need the following lemma that follows immediately from the above definitions. Lemma 4.1.5. [37, Lemme 4.13] Let (F, ∂F ) ⊂ (E(L), ∂E(L))) be an embedded surface in normal form with respect to the polyhedral decomposition and let γ1 , . . . , γk denote the components of ∂F . Then, k a(F ) ≥ j=1 c (γj ) . Finally we need the following. Lemma 4.1.6. Suppose that E(L) is an augmented link obtained from a prime, twist-reduced link diagram D(K). Let (F, ∂F ) ⊂ (E(L), ∂E(L))) be an embedded normal surface and let γ be a component of ∂F that is homologically non-trivial on a component T ⊂ ∂E(L). If T comes from a crossing circle of L, let m denote the number of crossings in the corresponding twist region of D(K). If T comes from a component Kj of K, let m be the number of twist regions visited by Kj , counted with multiplicity. c (s) ≥ mπ . 3 Proof. It is proven in the proof of [37, Corollary 5.12] using [37, Proposition 5.3]. 101 4.1.3 Genus estimates for spanning surfaces Let (S, ∂S) ⊂ (E(K), ∂E(K)) be a spanning surface of K. That is a surface where the components of ∂S are the components of K and it contains no closed components. In fact, often, we will use (some times implicitly) the following convenient characterization of spanning surfaces. Lemma 4.1.7. A properly embedded surface (S, ∂S) ⊂ (E(K), ∂E(K)), without closed components, is a spanning surface of K if and only if for each component of ∂E(K), the total geometric intersection number of the boundary curves ∂S with the corresponding meridian is 1. Proof. If S is a spanning surface of K, then clearly the desired conclusion holds. Conversely, suppose that each component of ∂S has geometric intersection number 1 with the meridian on the component of ∂E(L) it lies. Then ∂S must have exactly one component on each component of ∂E(K) . For, if we had more that one curves on some component of ∂E(K) then these curves will be parallel creating more intersections of ∂S with the corresponding meridian that would contribute to the geometric intersection number. Thus each component of ∂S is a longitude of ∂E(L). Let J be an augmented link obtained from a diagram of a K and let L be the corresponding fully augmented link. A spanning surface (S, ∂S) ⊂ (E(K), ∂E(K)) gives punctured surface (F, ∂F ) ⊂ (E(J), ∂E(J)). Now F gives a properly embedded surface (F , ∂F ) ⊂ (E(L), ∂E(L)). Since we are only interested in the Euler characteristic of the surface we will often choose to work with F instead of F . In fact, abusing the setting, we will identify F with F and say we can view F as a surface in the exterior of the fully 102 augmented link E(L) ∼ = E(J). For the next theorem we will also assume that the surface F can be isotoped into normal form with respect to the polyhedral decomposition of E(L). Theorem 4.1.8. Let K ⊂ S 3 be a link of k components with a prime, twist-reduced diagram D(K). Suppose that D(K) has t ≥ 2 twist regions and let τ denote the smallest number of crossings corresponding to a twist region of D(K). Let S be a spanning surface of K and let L be an augmented link, obtained from K, such that S intersects n crossing circles of L. Suppose that the corresponding punctured surface F ⊂ E(L) can be isotoped to be normal with respect to the polyhedral decomposition P1 , P2 . Then we have −χ(S) ≥ nτ t + −n . 3 6 Proof. The boundary ∂F consists of curves γ1 , . . . , γk one for each torus component of ∂E(L) that comes from a component of K and curves γk+1 , . . . , γk+n on the components coming from crossing circles. By assumption F can be isotoped into normal form in the polyhedra P1 and P2 ; so we can compute its combinatorial area. By applying Proposition 4.1.4 and Lemma 4.1.5 we have −2π · χ(S) = −2π · χ(F ) − 2πn = a(F ) − 2πn k ≥ i=1 n (γi ) + i=1 (γk+i ) − 2πn. By Lemma 4.1.6, the total length of the curves γ1 , . . . , γk is at least 2tπ/3, because K passes through each twist region twice. By the same lemma the total length of the curves 103 γk+1 , . . . , γk+n is at least nτ π/3. Thus from the last equation we obtain nτ π 2tπ + − 2πn 3 3 t nτ = 2π + −n . 3 6 −2π · χ(S) ≥ Since the Euler characteristic is an integer, the conclusion follows. Recall that for a non-orientable surface S, with k boundary components, the crosscap number is C(S) = 2 − χ(S) − k. We have the following result that should be compared with [37, Theorem 1.5]. Corollary 4.1.9. Let the notation and setting be as in Theorem 4.1.8. Suppose moreover that D(K) has at least six crossings in each twist region and that S is non-orientable. Then we have C(S) ≥ t 3 + 2 − k. Proof. Let τ denote the smallest number of crossings corresponding to a twist region of D(K). By assumption, τ ≥ 6. Thus Theorem 4.1.8 gives t nτ + − n +2 − k 3 6 t t + n − n +2 − k = + 2 − k. 3 3 C(S) = −χ(S) + 2 − k ≥ ≥ 104 4.2 Crosscap numbers of Alternating links Adams and Kindred [1] gave an algorithm, starting with an alternating link projection, to construct a spanning surface of maximal Euler characteristic among all the spanning surfaces of the link. Our goal in this section is to show that the surface obtained from the algorithm of [1] lies in the complement of an appropriate augmented link obtained from the link projection. In the next section we will use the techniques of Section 4.1 to estimate the crosscap number of a prime alternating link in terms of the twist number and the crossing number of any prime, twist-reduced alternating diagram of the link. 4.2.1 State surfaces and a minimum genus algorithm Given a crossing on a link diagram D(K) there are two ways to resolve it. A Kauffman state σ on D(K) is a choice of one of these two resolutions at each crossing of D(K). Given a state σ of D(K) we obtain a spanning surface Sσ of K, as follows: The result of applying σ to D(K) is a collection vσ (D) of non-intersecting circles in the plane, called state circles. We record the crossing resolutions along σ by embedded segments connecting the state circles. Each circle of vσ (D) bounds a disk in S 3 . These disks may be nested on the projection plane but can be made disjointly embedded in S 3 by pushing their interiors at different heights below the projection plane. For each arc recording the resolution of a crossing of D(K) in σ, we connect the pair of neighboring disks by a half-twisted band. The result is a surface Sσ ⊂ S 3 whose boundary is K. See Figure 4.5. Note that there might be several ways to make the disks bounded by the circles vσ (D) disjoint and the resulting surface may not necessarily be isotopic in S 3 . Nevertheless, this point is not important for the results of this paper, or for those of [1] as the resulting surfaces 105 circle of v (D) bounds a disk in S 3 . These disks may be nested on the projection plane but can be made disjointly embedded in S 3 by pushing their interiors at di↵erent heights below the projection plane. For each arc recording the resolution of a crossing of D(K) in , we connect the pair of neighboring disks by a half-twisted band. The result is a surface S ⇢ S 3 whose boundary is K. See Figure 5. Figure 5. The two resolutions of a crossing, the arcs recording them Figure 4.5: and Thetheir two contribution resolutions of crossing, the arcs recording them and their contribution to astate surfaces. to state surfaces. will have the same topology (i.e. Euler characteristic and orientability). The geometry of the state surfaces, Sσ with the particular construction described above, was studied by Futer, Kalfagianni and Purcell [34, 36]. In [1] the authors show that state surfaces of alternating diagrams can be used to determine the crosscap numbers of alternating links. In particular, starting with an alternating diagram D(K), they gave an algorithm to obtain a state surface with maximal Euler characteristic among all the state surfaces of D(K). Then they showed that this Euler characteristic is the maximum over all spanning surfaces of K, state or nonstate. To review this algorithm, recall that a diagram D(K) may be viewed as a 4-valent graph on S 2 with over and under information at each vertex. A complementary region of this graph on S 2 is an m-gon if its boundary consists of m vertices or edges. We will refer to a 2-gon as a bigon and a 3-gon as a triangle. We need the following elementary lemma. Lemma 4.2.1. Suppose that an alternating diagram D(K) contains no 1-gon or bigon. Then it must contain at least one triangle. Proof. Consider the 4-valent graph on S 2 defined by D(K) and let V , E, F denote the number of its vertices, edges and complimentary regions, respectively. Then, V − E + F = 2. We have E = 4V /2 = 2V . Hence V − 2V + F = 2, which implies that F > V . Let m be the smallest positive integer for which the 4-valent graph contains an m-gon. If it contains no bigon or triangle, then m > 3. This implies that F < 4V /4 = V since each face must have at least four distinct vertices in its boundary and each vertex can only be on the boundary 106 of at most 4 distinct faces. This is a contradiction. Therefore, m > 3 and if m > 2, then m = 3. Observe that the Euler characteristic of a state surface, corresponding to a state σ, is χ(Sσ ) = vσ − c, where c is the number of crossings on D(K). Thus to maximize χ(Sσ ) we must maximize the number of state circles vσ . Now we review the algorithm of [1]: Algorithm 4.2.2. Let D(K) be a connected, alternating diagram. 1. Find the smallest m for which the projection D(K) contains an m-gon. 2. (a) If m = 1, then resolve the corresponding crossing so the 1-gon becomes a state circle. Suppose that m = 2. Then D(K) contains twist regions with more than one crossings. Pick R to be such a twist region with cR > 1 crossings and cR − 1 bigons. Resolve all the crossings of R in such a way so that all these bigons become state circles. Create one branch of the algorithm for each bigon on D(K). (b) Suppose m > 2. Then by Lemma 4.2.1, we have m = 3. Pick a triangle region on D(K). Now the process has two branches: For one branch we resolve each crossing on this triangle’s boundary so that the triangle becomes a state circle. For the other branch, we resolve each of the crossings the opposite way. See Figure 4.6. 3. Repeat Steps 1 and 2 until each branch reaches a projection without crossings. Each branch corresponds to a Kauffman state of D(K) for which there is a corresponding state surface. Of all the branches involved in the process choose one that has the largest 107 CROSSCAP NUMBERS AND THE JONES POLYNOMIAL 11 Figure 6. One branch of the algorithm resolves the crossings so that Figure 4.6: One branchbecomes of the algorithm resolves the crossings so that thethe triangle the triangle a state circle. The other branch resolves them becomes a state circle. The other branch resolves them the opposite way. opposite way. the largest number corresponding of state circles. to The surface numberone of that statehas circles. The surface this state corresponding has maximal toEuler this state has maximal Euler characteristic over all the states corresponding to D(K). Note that, a priori , more than one branches of the algorithm may lead characteristic over all the states corresponding to D(K). Note that, a priori , more than to surfaces of maximal Euler characteristic. That is, there might be several state surfaces that have the same (maximal) Euler characteristic. one branches of the algorithm may lead to surfaces of maximal Euler characteristic. The following result allows to calculate the crosscap number of an alternating link from obtained applying Algorithm 3.2that to any alternating projection of the Thatthe is, surface there might bebyseveral state surfaces have the same (maximal) Euler link. characteristic. Theorem 3.3. [1, Corollary 6.1] Let S be any maximal Euler characteristic surface obtained via Algorithm 3.2 from an alternating diagram of k-component link K. Let C(K) denote the crosscap number of K and let g(K) denote the (orientable) genus of The K.following Then, weresult have allows to calculate the crosscap number of an alternating link from (1) If there is a surface S as above that is non-orientable then C(K) = 2 (S) k. the surface applying Algorithm to any we alternating projection of thek.link. (2)obtained If all theby surfaces S as above are4.2.2 orientable, have C(K) = 3 (S) Furthermore S is a minimal genus Seifert surface of K and C(K) = 2g(K) + 1. It should clarified that 6.1] choices as well asEuler the order in resolvingsurface bigon obTheorem 4.2.3.be[1, Corollary LetofS branches be any maximal characteristic regions following algorithm 3.2 may result in di↵erent state surfaces. If a state surface resulting from is orientable, it of is not necessarilylink the K. caseLet that tainedfor viaKAlgorithm 4.2.2 the fromalgorithm an alternating diagram k-component C(K) C(K) = 3 (S) k. If any state surface resulting from a choice of the branch and of bigon regions to resolve is non-orientable, then C(K) = 2 (S) k. We denoteorder the crosscap number of K and let g(K) denote the (orientable) genus of K. Then, illustrate the subtlety in the algorithm by applying it to the figure 8 knot. The figure 8 knot has non-orientable genus 2 realized by its checkerboard surfaces. we have In the first figure above, a diagram of the figure 8 knot is shown as well as two bigon regions of the diagram to which we can apply Step (a) of Algorithm 3.2. Suppose that choose region 1, then for the next step of the algorithm, we now have 3 choices of 1. we If there is a surface S as above that is non-orientable then C(K) = 2 − χ(S) − k. bigon regions to resolve. This is shown in the second figure above. If we choose region 1, then we obtain a checkerboard surface shown as a first figure below. If we choose 2 or 3, then Sweasobtain surface shown as the second figure.Furthermore Both of 2. region If all the surfaces above an areorientable orientable, we have C(K) = 3−χ(S)−k. these surfaces realize the maximal Euler characteristic of -1. It would be incorrect to conclude based on orientable surface that the non-orientable genus is 3 since a particular S is a minimal genus Seifert surface of K and C(K) = 2g(K) + 1. application of the algorithm gives a non-orientable surface. For our purposes, we will not need to know whether the resulting state surface from the algorithm is non-orientable therefore realize the non-orientable genus. Weasneed to know that maximizes It and should be clarified that choices of branches wellonly as the order in itresolving bigon the Euler characteristic. The next lemma is important for the results in this paper as it will allow us to apply regions following algorithm 4.2.2 may result in different state surfaces. If a state surface the techniques of Section 2 to obtain bounds on crosscap numbers of alternating links. for K resulting from the algorithm is orientable, it is not necessarily the case that C(K) = 108 3 − χ(S) − k. If any state surface resulting from a choice of the branch and order of bigon regions to resolve is non-orientable, then C(K) = 2 − χ(S) − k. We illustrate the subtlety in the algorithm by applying it to the figure 8 knot. The figure 8 knot has non-orientable genus 2 realized by its checkerboard surfaces. 1 3 1 2 2 Figure 4.7: A diagram of the figure 8 knot with bigon regions labeled 1 and 2 and the diagram resulting from applying the first step of Algorithm 4.2.2. The new choices of bigon regions are labeled 1, 2, and 3. In the first figure above, a diagram of the figure 8 knot is shown as well as two bigon regions of the diagram to which we can apply Step (a) of Algorithm 4.2.2. Suppose that we choose region 1, then for the next step of the algorithm, we now have 3 choices of bigon regions to resolve. This is shown in the second figure above. If we choose region 1, then we obtain a checkerboard surface shown as a first figure below. If we choose region 2 or 3, then we obtain an orientable surface shown as the second figure. Both of these surfaces realize the maximal Euler characteristic of -1. It would be incorrect to conclude based on orientable surface that the non-orientable genus is 3 since a particular application of the algorithm gives a non-orientable surface. For our purposes, we will not need to know whether the resulting state surface from the algorithm is non-orientable and therefore realize the non-orientable genus. We need only to know that it maximizes the Euler characteristic. 109 3 3 1 1 2 2 Figure 4.8: The state surfaces resulting from resolving the rest of the bigon regions with different choices of bigon regions. The next lemma is important as it will allow us to apply the techniques of Section 4.1 to obtain bounds on crosscap numbers of alternating links. Lemma 4.2.4. Given a prime, alternating, twist-reduced diagram D(K), let S be a maximal Euler characteristic surface obtained by applying Algorithm 4.2.2 to D(K). There is an augmented link J = JS such that S is in the complement E(J). Proof. An augmented link is obtained from D(K) by adding a simple closed curve encircling each twist region. If a twist region involves more than one crossing, then there is only one way to add a crossing circle. Otherwise, there are two ways of adding a crossing circle as shown in Figure 4.3. Let S be a state surface with maximal Euler characteristic obtained by applying Algorithm 4.2.2 to D(K). We show that we can always augment D(K) by making a choice of a crossing circle for each twist region involving a single crossing, such that S lies in the complement of the augmented link. For a twist region R involving more than one crossing, and thus consists of a maximum string of complementary bigon regions arranged end to end, we augment by adding a crossing circle CR encircling the twist region. Since D(K) is prime, none of its complementary regions can be an 1-gon. In this case, the algorithm picks the resolution of the crossings of R so 110 that each bigon becomes a state circle following Step 2a. In any state surface which have this resolution at the crossings of R, these bigon disks are joined with twisted strips, and the crossing circle CR encloses the twisted strips. We may arrange so that each twisted strip intersects the crossing disk corresponding to CR only in its interior. Hence, the portion of CROSSCAP NUMBERS AND THE JONES POLYNOMIAL 13 any state surface obtained by the algorithm involving a twist region containing at least one interior. Hence, the portion of any state surface obtained by the algorithm involving a twist at least one lie in 4.9. the complement of CR . See Figure 9. bigon, willregion lie incontaining the complement of Cbigon, . Seewill Figure R Figure 9. The portion of S through a twist region with more than one Figure 4.9: crossing The portion of crossing S through a twist with more than one crossing and the and the circle for theregion twist region. crossing circle for the twist region. Continue creating a branch of the algoritm for each bigon. Examine the link diagram D0 created after resolving all the twist regions involving more than one crossings by Algorithm 3.2. Continue creating a branch of the algorithm for each bigon. Examine the link diagram D Each of the remaining crossings corresponds to a twist region of D(K) containing a single crossing. We all decide on which way involving to add a crossing circle to crossings each of these twist created after resolving the twist regions more than one by Algorithm regions. Since D(K) is twist-reduced, bigons in D0 can only come from triangles in region with more than one crossings attached to them. If there 4.2.2.D(K) that had a twist is such a bigon in D0 , Algorithm 3.2 will apply Step 2a to these crossings such that the bigon becomes a state circle. If this bigon is adjacent to a nother one then we choose an Each of the remaining crossings corresponds to a twist region of D(K) containing a single augmentation for each of the two crossings by putting two crossing circles, one for each of the two crossings resolved, so that each crossing circle encloses a twisting strip from the crossing. We decide on whichThe wayportion to addof athe crossing circle coming to eachfrom of these twist regions. resolution at one crossing. state surface resolving these two crossings will then also be disjoint from the crossing circles. Repeat this procedure Sincefollowing D(K) isthe twist-reduced, bigons D nocan only come from triangles in D(K) that had a algorithm until thereinare more bigons. If there are no more crossings left in the projection, then we are done. Otherwise, we twisthave region with more one crossingsn-gon attached to them. If there3.1, is such a bigon a projection forthan whom a minimal is a triangle by Lemma and Step 2b ofin D , the algorithm is applied to resolve each of the three crossings of the triangle, see Figure 6. Algorithm 4.2.2 each will apply 2a tocrossings these crossings that a state In addition, of the Step remaining is a twistsuch region in the the bigon originalbecomes projection D(K). Let S be a maximal Euler characteristic surface obtained from the algorithm. timebigon that Step 2b is applied, a branch of thewe algorithm is chosen to generate circle.Each If this is adjacent to another one then choose an augmentation for S. each of We augment each of the three crossings of the triangle based on which branch is chosen. In either branch, we add two a crossing circle for each three crossings, such that it the two crossings by putting crossing circles, one of forthe each of the two crossings resolved, encircles the crossing strip from the resolution of the chosen branch. See Figure 10. so that each crossing circle encloses a twisting strip from the resolution at one crossing. The portion of the state surface coming from resolving these two crossings will then also be 111 Figure 10. Surfaces from two branches of the splitting in Step 2b in the Figure 9. The portion of S through a twist region with more than one crossing and the crossing circle for the twist region. the algoritm for each bigon. Examine the link diagram disjointContinue from thecreating crossinga branch circles. ofRepeat this procedure following the algorithm until there D0 created after resolving all the twist regions involving more than one crossings by Algorithm 3.2. are no more bigons. Each of the remaining crossings corresponds to a twist region of D(K) containing a We decide on which way to projection, add a crossing circle of these twist we Ifsingle therecrossing. are no more crossings left in the then we to areeach done. Otherwise, 0 regions. Since D(K) is twist-reduced, bigons in D can only come from triangles in that hadfor a twist with more than crossings to them. there2b of have D(K) a projection whomregion a minimal n-gon is aone triangle by attached Lemma 4.2.1, andIf Step is such a bigon in D0 , Algorithm 3.2 will apply Step 2a to these crossings such that the bigon becomes a state circle. If this bigon is adjacent to a nother one then we choose an the algorithm is applied to resolve each of the three crossings of the triangle, see Figure 4.6. augmentation for each of the two crossings by putting two crossing circles, one for each of the two crossings resolved, so that each crossing circle encloses a twisting strip from the In addition, each of the remaining a twist the original projection D(K). resolution at one crossing. The crossings portion ofisthe state region surface in coming from resolving these two crossings will then also be disjoint from the crossing circles. Repeat this procedure Let Sfollowing be a maximal Euler until characteristic surface obtained from the algorithm. Each time the algorithm there are no more bigons. If there are no more crossings left in the projection, then we are done. Otherwise, we that have Step a2bprojection is applied, branch of the algorithm chosenby to Lemma generate S.and WeStep augment forawhom a minimal n-gon is aistriangle 3.1, 2b of each the algorithm is applied to resolve each of the three crossings of the triangle, see Figure 6. of theInthree crossings of the theremaining triangle based on which branch is in chosen. In either branch, we addition, each of crossings is a twist region the original projection D(K). Let S be a maximal Euler characteristic surface obtained from the algorithm. time circle that Step 2b is of applied, a branch of thesuch algorithm chosen tothe generate S. strip add aEach crossing for each the three crossings, that itisencircles crossing We augment each of the three crossings of the triangle based on which branch is chosen. branch,ofwe a crossing circle each of the three crossings, such that it from In theeither resolution theadd chosen branch. Seefor Figure 4.10. encircles the crossing strip from the resolution of the chosen branch. See Figure 10. Figure 10. Surfaces from two branches of the splitting in Step 2b in the and theofcorresponding of augmentation. Figure 4.10:Adam-Kindred Surfaces fromalgorithm two branches the splitting choice in Step 2b in the Adam-Kindred algorithm and the corresponding choice of augmentation. We repeat as in Step 3 of Algorithm 4.2.2 which will make a decision for splitting at the rest of the crossings. If a bigon is encountered again as a minimal n-gon we repeat the procedure for a bigon region where each of the crossings belongs to a twist region in D(K). Otherwise we apply the procedure for when a minimal n-gon is a triangle. We stop when there are no more twist regions in D(K) to augment. 112 4.2.2 Normalization and crosscap number estimates Here we use the results of Section 4.1 to obtain two-sided bounds of the crosscap number of an alternating link in terms of the twist number of any prime, twist-reduced alternating link diagram. We begin with the following lemma that shows that the crosscap number of an alternating link is always bounded by the twist number of any alternating projection from above; that is, primeness and twist-reducibility is not needed for this part. Lemma 4.2.5. Let K ⊂ S 3 be a link of k components with an alternating diagram D(K) that has t ≥ 2 twist regions. Let C(K) denote the crosscap number of K. We have C(K) ≤ t + 2 − k. Proof. Let S be a surface obtained by applying Algorithm 4.2.2 to D(K) and let σ denote the Kaufmann state of D(K) to which S corresponds. Let vb denote the number of state circles that are bigons and let vnb denote the non-bigon state circles. Also let c1 denote the number of crossings in D(K) each of which forms its own twist region and let c2 denote the remaining crossings. Since we assume that there are at least two twist regions we have vnb ≥ 1. Thus we have −χ(S) = = c2 − vb + c1 − vnb ≤ t − 1. 113 Now the upper bound follows at once from Theorem 4.2.3. Before we are able to estimate C(K) from below we need some preparation. Let D(K) be a link diagram and let J be an augmented link obtained from D(K) with L the corresponding fully augmented link. Suppose that the exterior E(J) ∼ = E(L) is hyperbolic with an angled polyhedral decomposition {P1 , P2 } as described in Section 4.1. Let S be a spanning surface of K that realizes C(K). Having Lemma 4.2.4 in mind, we will also assume that S is disjoint from the crossing circles of J; hence we can view it as a surface in E(J) ∼ = E(L). We wish to apply Theorem 4.1.8 to estimate χ(S). In order to do so we need to have S in normal position with respect to the polyhedral decomposition {P1 , P2 }. The standard argument of making an surface that is incompressible and ∂-incompressible (that is, essential ) normal in a triangulated 3-manifold [50], can be adjusted to work in the setting of more general polyhedral decompositions. The argument is written down by Futer and Guritaud [30, Theorem 2.8]. An additional difficulty in our setting is that surfaces that realize C(K) need not be essential in E(K). For instance, if all the state surfaces obtained from Algorithm 4.2.2 applied to an alternating projection D(K) are orientable, then a spanning surface that realizes C(K) is a obtained from a minimal genus Seifert surface by adding a half-twisted band. Such a surface is ∂-compressible. This, for example, happens for the knot 74 [1]. For our purposes we are only interested in the question of whether S can be converted to a spanning surface that is normal with respect to the polyhedral decomposition of E(L), without changing χ(S) and the surface orientability. For this we examine how a spanning surface S that realizes C(K) behaves under the general process that converts any properly embedded surface in E(J) into a normal one with possibly different topology [66]. We 114 have the following lemma that applies beyond the class of alternating links and might be of independent interest. Lemma 4.2.6. Let K ⊂ S 3 be a link with a prime, twist-reduced diagram D(K). Suppose that D(K) has t ≥ 2 twist regions. Suppose that there is a spanning surface S in the exterior E(J) ∼ = E(L) of an augmented link of D(K) and such that C(S) = C(K). Then exactly one of the following is true: 1. There is a non-orientable, spanning surface S ⊂ E(L) for K, that is in normal form and such that χ(S) = χ(S ). 2. We have C(K) = 2g(K) + 1. Furthermore, there a Seifert surface of K that lies in E(L) such that it realizes g(K) and it is in normal form. Proof. By assumption S realizes C(K). Thus S has maximal Euler characteristic among all non-oriented spanning surfaces of K. As discussed above, we will assume that S is not necessarily essential and examine how compressions and ∂-compressions may interfere with a process of converting S to a normal surface. Examining the moves required during this process [30, Theorem 2.8], and since S contains no closed components, we see that there are two situations to consider: 1. S admits a compression disk D, that lies in the interior of a single face of a polyhedron P ∈ {P1 , P2 }. 2. The intersection of S with a face of a polyhedron f ⊂ ∂P is an arc γ that runs from an edge of e ⊂ Γ ⊂ ∂P to itself, or from an interior edge to an adjacent boundary edge. In (1) there are two cases to consider according to whether ∂D separates S or not. First suppose that ∂D is non-separating on S. Compressing along D we obtain a spanning surface 115 S for K with χ(S ) = χ(S) + 2. Since S realizes C(K), S cannot be non-orientable. Thus we have an orientable spanning surface of K; that is a Seifert surface. Adding a half-twisted band to S (i.e. adding a crosscap) produces a non-orientable spanning surface S1 of K with χ(S1 ) = χ(S ) − 1. Thus, S1 is a non-orientable spanning surface with χ(S1 ) = χ(S) + 1 > χ(S) which contradicts the fact that S realizes C(K). Thus this case will not happen. Suppose now that ∂D separates S. We will look at such a disk so that ∂D is innermost in the sense that one of the components of S ∂D lies entirely in a single polyhedron P . Compressing along D gives two surfaces S1 and S2 with χ(S) = χ(S1 ) + χ(S2 ) − 2. Suppose that both surfaces have non-empty boundary. Then the disjoint union of S1 , S2 is a non-orientable spanning surface of K with Euler characteristic χ(S1 ) + χ(S2 ) > χ(S), contradicting the assumption that S realizes C(K). Thus, one of S1 , S2 , say S1 , must be a closed surface and all of ∂S is left on S2 . Since S1 is a closed surface embedded in S 3 , it is orientable. Hence S2 is a non-orientable. Since S realizes C(K) and χ(S1 ) ≤ 2, it follows that χ(S) = χ(S2 ). Thus we may ignore S2 , replace S with S2 and continue with the normalization process. Next we treat case (2): The arc γ cuts off a disk D ⊂ f , with ∂D consisting of γ and an arc that lies on the boundary of f . By an innermost argument, we may assume that the interior of D is disjoint from S. Again, following the argument in the proof of [30, Theorem 2.8], if γ runs from an interior edge to an adjacent boundary edge, the disk D will guide an isotopy of S that can be used to eliminate the arc γ [30, Figure 2.2]. Similarly, if e is an interior edge or e is a boundary edge and f is a boundary face, we can use D to obtain an isotopy that will eliminate γ and decrease the number intersections of S with Γ (compare, left panel of [30, Figure 2.1]). It follows, that the only case left to examine is when e is 116 a boundary edge and f is an interior face of S. In this case D is a ∂-compression disk of S. Consider the arc δ := ∂D γ ⊂ e and let T denote the boundary component of ∂E(K) containing it. There are two cases to consider: (i) δ cuts off a disk in the annulus T ∂S; and (ii) δ runs from one component of T ∂S. We will perform surgery (∂-compression) along D. This may cut S into more components or not according to whether we are in case (i) or (ii). Surgery along a ∂D doesn’t change the total algebraic intersection number of ∂S with the meridians of ∂E(J). Thus, by Lemma 4.1.7, it will produce a spanning surface of K and possibly some (redundant) components. First suppose that surgery along D splits S into two surfaces S1 and S2 . Suppose that both surfaces have non-empty boundary. Then the disjoint union of S1 , S2 is a non-orientable spanning surface of K with Euler characteristic χ(S1 ) + χ(S2 ) > χ(S), contradicting the assumption that S realizes C(K). Thus all the components of ∂S disjoint from D must remain on one of S1 , S2 , say on S2 and the intersections of ∂S with the meridians of ∂E(K), that count towards algebraic intersection number, also remain on ∂S2 . Thus S1 has one boundary component that is either homotopicaly trivial on ∂E(K) or isotopic to a meridian of ∂E(K). In either case ∂S1 bounds a disk in S 3 . We may cap ∂S1 with this disk to produce a closed surface embedded in S 3 , which must be orientable. Hence S2 is a non-orientable spanning surface for K. But since S realizes C(K) and χ(S1 ) ≤ 1 we must have χ(S) = χ(S2 ). Thus we may ignore S1 , replace S with S2 and continue with the normalization process. Next suppose that surgery along D doesn’t disconnect S. Then we get a spanning surface S , with χ(S ) = χ(S)+1. Since S was assumed to realize C(K), S cannot be non-orientable. Thus we have an orientable spanning surface of K. We claim that S must be a minimal genus Seifert surface, that is g(S ) = g(K). For, suppose that K has a Seifert surface S1 with 117 χ(S1 ) > χ(S ). Then adding a half-twisted band to S1 would give a non-oriented spanning surface S with χ(S ) = χ(S1 ) − 1 > χ(S ) − 1 = χ(S), contradicting the fact that S realizes C(K). Thus, S is a minimal genus Seifert surface of K that lies in E(L); that is we have g(S ) = g(K). Now it is clear that the surface obtained by a half-twisted band to S is a nonorientable spanning surface S of K that has maximal Euler characteristic among all such surfaces. Thus we have C(K) = C(S ) = 2g(S ) + 1 = 2g(K) + 1. Since S is minimal genus and orientable, it is incompressible and ∂-incompressible in E(K) and thus in E(L). Hence we may isotope S to be normal with respect to the polyhedral decomposition [30, Theorem 2.8]. Now we are ready to prove the following. Theorem 4.2.7. Let K ⊂ S 3 be a link of k components with a prime, twist-reduced alternating diagram D(K). Suppose that D(K) has t ≥ 2 twist regions. Let C(K) denote the crosscap number of K. We have t 3 + 2 − k ≤ C(K) ≤ t + 2 − k, where · is the ceiling function that rounds up to the nearest integer. Furthermore, both bounds are sharp. Proof. The upper bound comes from Lemma 4.2.5. To derive the lower bound, let S be a state surface obtained from Algorithm 4.2.2 to D(K) and let J be the augmented link of Lemma 4.2.4, with L the corresponding fully augmented link. By Theorem 4.2.3, one of following is true: 1. There is one such surface that is non-orientable and realizes C(K); that is we have C(K) = C(S). 118 2. All such surfaces S are orientable and we have C(K) = 2g(S) + 1. NUMBERS AND THE JONES POLYNOMIAL 17 Suppose we are in caseCROSSCAP (1). Then by Lemma 4.2.6 we may replace S by a spanning surface ⇠ ⇡ that also realizes C(K) and is normal with respect to the polyhedral decomposition of E(L). t + 2 k  C(K)  t + 2 k, 3 Theorem where4.1.6 d·e isgives the ceiling function that rounds up to the nearest integer. Furthermore, both bounds are sharp. Proof. The upper bound comes from Lemma 3.5. To derive the lower bound, let S be t a state surface obtained Algorithm 3.2 to − D(K) k, augmented link C(K) =from C(S) = 2 − χ(S) k ≥and let+J 2be−the 3 of Lemma 3.4, with L the corresponding fully augmented link. By Theorem 3.3, one of following is true: (1) There is one such surface that is non-orientable and realizes C(K); that is we and the lower bound have C(K)follows. = C(S).On the other, hand if we are in (2) then S is a minimal genus (2) All such surfaces S are orientable and we have C(K) = 2g(S) + 1. Seifert surfaceweofare K in and thus is incompressible ∂-incompressible. Thus wesurface may again Suppose case (1).itThen by Lemma 3.6ands we may replace S by a spanning that also realizes C(K) and is normal with respect to the polyhedral decomposition of replace S with one into normal form. Since S is disjoint from the crossing circle components E(L). Theorem 2.6 gives ⇠ ⇡ t C(K) = C(S) = 2 (S) k + 2 k, of J, Theorem 4.1.6 gives 3 and the lower bound follows. On the other, hand if we are in (2) then S is a minimal genus Seifert surface of K and thus it is incompressible ands @-incompressible. Thus we may again replace one+into form. S is tdisjoint from C(K)S=with 2g(S) 1 = normal 2 − 2χ(S) − kSince +1≥ +3− k. the crossing 3 circle components of J, Theorem 2.6 gives ⇠ ⇡ t C(K) = 2g(S) + 1 = 2 2 (S) k + 1 +3 k. 3 Hence the lower bound follows. Hence the lower bound follows. Figure 11. The knots 103 (left) and 10123 (right). Figure 4.11: The knots 103 (left) and 10123 (right) [19]. It remains to prove that both bounds are sharp: Consider the alternating knot 103 of Figure 11. The twist number of the diagram shown there is t = 2 and each twist consists of more than one bounds crossing.are Thus Algorithm 3.2 has one branch that103 of Itregion remains to prove that both sharp: Consider theonly alternating knot is easily seen to give an oriented surface of genus 1. Thus, by Theorem 3.3, C(K) = 2g(K) + 1The = 3 twist = t + 1. The same argument applies to the knots adding anyregion Figure 4.11. number of the diagram shown there is t obtained = 2 andbyeach twist even number of crossings in each twist region of the knot 103 . Hence we have an infinite family of alternating knots with C(K) = 2g(K) + 1 = 3 = t + 1. consists of more than one crossing. Thus Algorithm 4.2.2 has only one branch that is easily To discuss some examples where our lower bound is sharp, note that if a knot K processes an alternating diagram with c = t, then we have seen to give an oriented surface of genus 1. Thus, by Theorem 4.2.3, C(K) = 2g(K) + 1 = lcm jc k 1+  C(K)  . 3 2 119 3 = t + 1. The same argument applies to the knots obtained by adding any even number of crossings in each twist region of the knot 103 . Hence we have an infinite family of alternating knots with C(K) = 2g(K) + 1 = 3 = t + 1. To discuss some examples where our lower bound is sharp, note that if a knot K processes an alternating diagram with c = t, then we have 1+ c 3 ≤ C(K) ≤ c . 2 Now observe that, for instance, if c = t = 13, then C(K) = 6. Similarly, if c = t = 10, then C(K) = 5. A concrete example, is the knot 10123 shown in Figure 4.11. More examples where the lower bound is sharp are discussed in Section 4.4. In particular, for knots we have the following: Theorem 4.2.8. Let K ⊂ S 3 be a knot with a prime, twist-reduced alternating diagram D(K). Suppose that D(K) has t ≥ 2 twist regions and let C(K) denote the crosscap number of K. We have 1+ t 3 ≤ C(K) ≤ min t + 1, c 2 where c denotes the number of crossings of D. Furthermore, both bounds are sharp. Proof. For k = 1, the inequality of Theorem 4.2.7 becomes t 3 + 1 ≤ C(K) ≤ t + 1. Thus the lower bound follows. Murakami and Yasuhara [89] showed that for a knot K 120 with a connected, prime diagram of c crossings, we have C(K) ≤ c . 2 Thus the upper bound follows from these two inequalities. Having related C(K) to twist numbers of alternating link projections, Theorems 1.1.3 and 4.0.1 follow by [26]. 4.3 Proof of Theorem 1.1.3 Now we discuss how the results stated in the introduction follow from the above results. Recall that for a knot K, we have TK := |βK | + βK , where βK and βK denote the second and the penultimate coefficients of the Jones polynomial of K, respectively. Theorem 1.1.3. Let K be a non-split, prime alternating link with k-components and with crosscap number C(K). Suppose that K is not a (2, p) torus link. We have TK 3 + 2 − k ≤ C(K) ≤ TK + 2 − k. Furthermore, both bounds are sharp. Proof. Let D(K) be a connected, twist-reduced alternating diagram that has t ≥ 2 twist regions. Then, by [26, Theorem 5.1], we have t = TK . A prime alternating link admits prime, twist reduced alternating diagrams; every alternating diagram can be converted to a twist-reduced one by flype moves [76, Lemma 4]. Thus, the result follows from Theorem 4.2.7. 121 Now we are ready prove 4.0.1 which we restate for the convenience of the reader. Theorem 4.0.1. Let K be an alternating, non-torus knot with crosscap number C(K) and let TK be as above. We have 1+ TK 3 ≤ C(K) ≤ min TK + 1, sK 2 where sK denotes the span of JK (t). Furthermore, both the upper and lower bounds are sharp. Proof. Kauffman [72] showed that the degree span of the Jones polynomial of alternating link is equal to the crossing number of the link. Furthermore, both the degree span and the crossing number of alternating knots are known to be additive under the operation of connect sum [78]. Using these, the upper inequality follows at once from Theorem 4.2.8. Furthermore, the lower inequality follows for all prime alternating knots. To finish the proof we need to show that the lower inequality holds for connect sums of alternating knots. To that end let K#K be such a knot. By [89], we have C(K#K ) ≥ C(K) + C(K ) − 1. Since the Jones polynomial is multiplicative under connect sum we have TK +T K Hence we have T T C(K#K ) ≥ C(K) + C(K ) − 1 = K + K + 4 − 2 − 1 3 3 TK + T K + 1. ≥ 3 122 ≥T K#K . Hence the conclusion follows. 4.4 4.4.1 Calculations of crosscap numbers Lower exact bounds The lower bound of Theorem 4.2.7 gives the exact value of the crosscap number in the case we have an alternating projection D(K), where each twist region of D(K) has at least two crossings and the twist regions of D(K) meet in groups of three. A concrete construction of such examples is as follows: Let G be a trivalent planar graph and let N (G) denote a neighborhood of G on the plane. The boundary ∂N (G) is a link. For each edge of G we have two parallel arcs belonging on different components of ∂N (G). Construct a diagram of a new link by adding a number of half twists between these parallel arcs of the components ∂N (G) corresponding to each edge of G. Suppose that for each edge of G we add at least two crossings on D(K). We can do this so that the resulting diagram is alternating. Depending on the numbers of twists we put we may obtain a knot or a multi-component link. Let D(K) denote any alternating projection obtained this way and let SG be a state surface obtained from Algorithm 4.2.2 applied to D(K): Since each twist region contains bigons, SG corresponds to the Kauffman state of D(K) that resolves all the crossings so that the bigons are state circles. To analyze these surfaces further we need a definition. Definition 4.4.1. A normal disk D in a polyhedron P ∈ {P1 , P2 } is called an ideal triangle if ∂D intersects exactly three boundary faces of ∂E(L) and it intersects no interior edges of P. 123 Corollary 4.4.2. Let K be a k-component link with a prime, twist-reduced alternating diagram D(K) constructed from a trivalent planar graph G as above. Then we have C(K) = where t 3 + = 2 if SG is non-orientable and −k = TK 3 + − k. = 3 otherwise. Proof. By assumption D(K) is obtained by adding twists along components of the boundary ∂N (G) of a regular neighborhood of a planar trivalent graph. The surface SG is obtained by N (G) by similar twisting and the augmented link of Lemma 4.2.4 is obtained by adding a component encircling each twist region of D(K). In order to calculate χ(SG ) we need some more detailed information about the polyhedral decomposition {P1 , P2 } of E(L) [98]. The surface SG gives rise to surface SG in E(L); we will calculate χ(SG ). Let D denote the union of the crossing disks bounded by the crossing circles of L. Each disk intersects the projection plane in a single arc. We may isotope the interior of SG so that it is disjoint from the intersections of D with the projection plane. Let P ∈ {P1 , P2 }. Recall that (before truncation) all the vertices of P are of valence four and they correspond to crossing circles of L and to arcs of K D and the faces can be colored in a checkerboard fashion (in shaded and white) as follows: 1. The shaded faces of P correspond to the crossing disks: Each disk D gives to two triangular shaded faces of P meeting at an ideal vertex corresponding to the crossing circle ∂D (“bowties”). 2. The edges of P come from the intersections of D with the projection plane. 3. The white faces of P correspond to regions of D(K) on the projection plane. 124 To a vertex v ∈ G there corresponds a triangular region of D(K) that is a neighborhood of v and around which the three twist regions D(K), corresponding to the edges of G emanating from v, meet. Let Dv denote the union of the three crossing disks corresponding to the three twist regions of D(K) around v. This triangular region will become a an ideal triangular white face, say Dv , after truncating the vertices of P . See Figure 4.12. The ideal vertices of Dv come from the arcs of K Dv that surround v. Each of the three interior edges of P on ∂Dv is attached to a bowtie: two shaded faces meeting at an ideal vertex coming from one of crossing disks of Dv . G Figure 4.12: From left to right: The portion of D(K) around a vertex of G, the corresponding potion of the augmented link and portion of the polyhedral decomposition. An ideal triangle is indicated by the red line. Since the surface SG doesn’t intersect the crossing circles, SG is disjoint from the ideal vertices corresponding to disks in Dv . Furthermore, since we arranged so that the interior of SG is disjoint from the intersection of D with the projection plane, SG is disjoint from the boundary faces of P corresponding to Dv . It follows that after putting SG into normal form in P we will have a normal triangle (an ideal triangle surrounding Dv ) corresponding to the part of SG around v. Now SG will consist of a collection of ideal triangles and the only contributions to a(SG ) will come from these ideal triangles. Let T be such a triangle 125 and let γi , i = 1, 2, 3, denote the three arcs of ∂T on ∂E(L). By Definition 4.1.3 and the definition of combinatorial lengths we have a(T ) = π = 3 · π = Σ3i=1 l(γi , T ). 3 The number of ideal triangles in Sg is at most 2tπ 3 . Hence the calculation in the proof of Theorem 4.2.7 gives a(SG ) = 2tπ , 3 and by Theorem 4.1.4 we obtain −χ(SG ) = −χ(SG ) = 3t . Now if SG is is non-orientable then, by Theorem 4.2.3, C(K) = C(SG ) = −χ(S) + 2 − k. If SG is orientable then C(K) = 2g(K) + 1 = −χ(SG ) + 3 − k. In both cases the desired result follows, since D(K) is prime and twist reduced and we have t = TK . Example 4.4.3. Let D(K) = P (p1 , p2 , . . . , pN ) denote the standard diagram of an alternating N -string pretzel knot. Suppose that, for i = 1, . . . , N , |pi | > 1. We can see that the surface S corresponding to Theorem 4.2.3 is the pretzel surface consisting of two disks and N -twisted bands; one for each twist region of D(K). Augment the diagram D(K) by adding a crossing circle at each twist region so that the surface S intersects each crossing disk in a single arc only. This surface is disjoint from the crossing circles of the fully augmented link L. Furthermore, the surface S is essential in E(K) and thus in E(L); see argument in the proof of Theorem 4.5.2 below. Hence we may put it in normal form to calculate the area a(S). Again by inspecting the combinatorics of the polyhedra decomposition as we did in the proof of Corollary 4.4.2, that the contributions to a(S) come from two identical, normal N -gons, D1 , D2 such that ∂Di intersects N -boundary faces of the polyhedral decomposition 126 and it intersects no interior edges. We have a(S) = a(Di ) + a(D2 ) = 2πN − 4π = 2π(TK − 2), and thus −χ(S) = N −2. If all, but one, pi are odd the S is non-orientable and thus C(K) = N − 1 = TK − 1. If all the pi ’s are odd, then S is orientable and then, C(K) = N = TK . Note that the crosscap numbers of pretzel knots have been calculated in [64]. 4.4.2 Low crossing knots Next we turn our attention to knots with small crossing numbers: The crosscap numbers of all alternating links up to 9 crossings are known. KnotInfo [19] provides an upper and a lower bound for the crosscap numbers of all knots up to 12 crossings for which the exact values of crosscap numbers are not known. Note that in all, but a handful of cases, where the crosscap number is not determined, the lower bound given in KnotInfo is 2. There are 1778 prime, alternating knots with crossing numbers 10 ≤ c ≤ 12. For these knots we calculated the quantity TK := |βK | + βK using the Jones polynomial value given in KnotInfo and we compared our crosscap number lower bound with the one given in therein. For 1472 of these knots our lower bound is better than the one given in KnotInfo and for 262 of them our lower bound agrees with the upper bound in there; thus we are able to calculate the exact value of the crosscap number in these cases. The remaining 306 knots are those where the exact crosscap number values are given in [19] or our lower bound is the same as the one in there. The computer file of our calculations can be found at http://www.math.msu.edu/∼kalfagia/crosscapdata.txt. For example, in Table 4.1 we have all the 37 alternating knots K for which KnotInfo 127 K 1085 11a97 11a263 11a323 12a0636 12a0845 12a1031 12a1142 12a1220 12a1285 TK 6 5 4 6 5 5 5 5 6 4 K 1093 11a223 11a279 11a330 12a0641 12a0970 12a1095 12a1171 12a1240 - TK 6 5 6 6 4 6 6 6 6 - K 10100 11a250 11a293 11a338 12a0753 12a0984 12a1107 12a1179 12a1243 - TK 6 5 6 4 5 6 6 6 4 - K 11a74 11a259 11a313 11a346 12a0827 12a1017 12a1114 12a1205 12a1247 - TK 5 5 6 6 5 6 6 6 6 - Table 4.1: Knots where the KnotInfo upper bound agrees with our lower bound. The crosscap number is 3. states 2 ≤ C(K) ≤ 3, together with the value of the corresponding quantity TK . In all the 37 cases the lower bound above is also 3; thus for all these knots we can determine the crosscap number to be 3. 4.5 4.5.1 Generalizations and questions Non-alternating links A question arising from this work is the question of the extent to which the Jones polynomial (coarsely) determines the crosscap number outside the class of alternating links. This is an interesting question that merits further investigation. Our contribution towards an answer to this question is to provide a generalization of Theorem 1.1.3 for some class of non-alternating links. To state our result we need a definition. Definition 4.5.1. For a link diagram D(K) let SA denote the state surface corresponding to the Kauffman state where all the crossings are resolved one way and let SB denote the 128 state surface corresponding to the state where all crossings are resolved the opposite way. A link K is called adequate if it admits a link diagram D(K) such that none of SA , SB contains a half-twisted band with both ends attached on the same state circle. Adequate links form a large class that contains the alternating ones but it is much wider. See [32, 34, 79]. Theorem 4.5.2. Let K be a k-component link with crosscap number C(K) and let TK be as above. Suppose that K admits a connected, twist-reduced, diagram D(K) that has t ≥ 2 twist regions, and such that each twist region of D(K) contains at least six crossings. We have t 3 + 2 − k ≤ C(K) ≤ t + 2 − k. If moreover D(K) is adequate then we have TK 6 + 2 − k ≤ C(K) ≤ 3TK − k − 1. Proof. Let S be a spanning surface of K that is of maximal Euler characteristic over all spanning surfaces (that is both oriented and non-oriented). Then S gives gives an incompressible and ∂-incompressible surface in E(K). For, surgery of S along a compression or a ∂-compression disk will produce a surface of higher Euler characteristic (compare proof of Lemma 4.2.6). Let L be a fully augmented link obtained from D(K). Recall that each crossing circle added in this process bounds a disk D whose interior is pierced exactly twice by K. We will isotope S in the complement of K so that S ∩ D is minimized. In this process, since S is incompressible, if there is a simple closed curve in δ ⊂ S ∩ D that bounds a disk in D with its interior disjoint from S, then we can eliminate δ by isotopy of S in the 129 complement of L. Similarly we can eliminate arc components of S ∩ D that cut off disks on D with their interior disjoint from S. Finally, simple closed curves that are parallel to ∂D can be eliminated by sliding S off the boundary of D. The surface S gives rise to a surface F in E(L). We claim that F is incompressible and ∂-incompressible in E(L). To see that, suppose that there is an essential simple closed curve γ ⊂ S that bounds a compressing disk in ∆ ⊂ E(L). Since S is incompressible, γ must bound a disk ∆ ⊂ S whose interior is intersected by the crossing circles of L. Now ∆ ∪ ∆ bounds a 3-ball that can be used to produce an isotopy that reduces the intersection of S and the crossing disks of L; contradiction. Now we argue that F is ∂-incompressible. To that end, suppose that F admits a ∂compression disk ∆, with ∂∆ = γ ∪ δ, where γ is a spanning arc in F and δ is an arc on a component T ⊂ ∂E(L). Assume, for a moment, that T corresponds to a component of K. Since S is ∂-incompressible in E(K), the arc γ must cut a disk ∆ ⊂ S whose interior is pieced by the crossing circles of L. Since S is incompressible, the boundary of the disk ∆ ∪ ∆ also bounds a disk ∆ ⊂ S. Now ∆ ∪ ∆ ∪ ∆ bounds a 3-ball that can be used to produce an isotopy that reduces the intersections of S with the crossing disks of L. This is a contradiction. Suppose now that T is a component of ∂E(L) that corresponds to a crossing circle of L. Since S intersects crossing circles an even number of times, T ∂S has at least two components. Thus δ lies on an annulus of A ⊂ T ∂S and either it cuts off a disk on A or it runs between different components of A. Now the usual argument that shows that an orientable, spanning surface of a link that is incompressible, has to be ∂-incompressible applies to obtain a contradiction (see [53, Lemma 1.10]). Thus the punctured surface S is essential in E(L). 130 Now we may replace S by surface, of the same orientability and Euler characteristic, that is in normal form with respect to the polyhedral decomposition of E(L) [30, Theorem 2.8]. Now Corollary 4.1.9 applies. If S is non-orientable, we have C(S) = C(K) and by Corollary 4.1.9 we have C(K) ≥ t 3 + 2 − k. If S is orientable then C(K) = 2g(S) + 1 = 2g(K) + 1 and by Theorem 4.1.8 again we have C(K) ≥ 3t + 2 − k. Combining these inequalities with Lemma 4.2.5 we have t 3 + 2 − k ≤ C(K) ≤ t + 2 − k, which proves the first part of the theorem. Suppose now that D(K) is also adequate. Then [32, Theorem 1.5] implies that t 3 ≤ TK ≤ 2t. Now combining the last two inequalities gives the desired result. Theorem 4.0.1 should be compared with Murasugi’s classical result [90] that the Alexander polynomial determines the Seifert genus of alternating knots. The Alexander polynomial doesn’t determine the genus of non-alternating knots. However the Knot Floer Homology (the categorification of the Alexander polynomial) determines the genus of all knots [94]. A related question is the question of whether the Khovanov homology [73] of knots (the categorification of the Jones polynomial) is related to the crosscap number and the extent to which the former determines the later. We close the subsection with the following questions: 131 Question 4.1. For which links does the Jones polynomial (coarsely) determine the crosscap number? Question 4.2. Are there two sided bounds of C(K) of every link K in terms of the Khovanov homology of K? 2 (q) and β be the For an adequate knot K, let β2 be the coefficient of q d(2) + 1 of JK 2 ∗ 2 (q). Recall that these are stable coefficients of the colored Jones coefficient of q d (2)−1 of JK polynomial, see Section 2.1.2 for details. The proof of Theorem 1.1.3 as well as Corollary β +|βK | TK = K 2π is related, and is often 4.4.2 and Example 4.4.3 show that the quantity 2π equal, to the combinatorial areas of a normal surfaces in augmented links of K. In view of this, one can also ask Question 4.3. Do the higher order stable coefficients of the colored Jones polynomial have similar interpretations in terms of topological quantities in augmented links from K? 132 Chapter 5 The Jones polynomial, 3-braids, and L-space knots An L-space knot generalizes the notion of knots which admit lens space surgeries. Recall that a rational homology 3-sphere Y is an L-space if |H1 (Y ; Z)| = rank HF (Y ), where HF denotes the ‘hat’ version of Heegaard Floer homology, and the name stems from the fact that lens spaces are L-spaces. Besides lens spaces, examples of L-spaces include all connected sums of manifolds with elliptic geometry [95]. A knot, K ⊂ S 3 , is an L-space knot if K or its mirror image admits a positive L-space surgery. Ozsv´ath and Szab´o’s result states that L-spaces admit no co-orientable taut foliations [94, Theorem 1.4]. It is also known that an L-space knot K ⊂ S 3 must be prime [74] and fibered [92], and that K supports the tight contact structure on S 3 [59, Proposition 2.1]. In addition, the Alexander polynomial K (t) of an L-space knot K satisfies the following: • The absolute value of a nonzero coefficient of K (t) is 1. The set of nonzero coefficients alternates in sign [95, Corollary 1.3]. • If g is the maximum degree of K (t) in t, then the coefficients of the term tg−1 is nonzero and therefore ±1 [60]. In this chaper, we study which 3-braids, that close to form a knot, admit L-space surgeries. We prove that: 133 Theorem 1.1.4. Twisted (3, q) torus knots are the only knots with 3-braid representations that admit L-space surgeries. We begin by computing certain coefficients of the Jones polynomials of closed 3-braids in Section 5.1. The Alexander polynomial of a closed 3-braid may be written in terms of the Jones polynomial [14]. This allows us to use our computation of the Jones polynomials to rule out closed 3-braids whose Alexander polynomials violate the aforementioned conditions. We state definitions and results from [31] on 3-braids that we need, and we use them to prove Theorem 1.1.4 in Section 5.2. 5.1 The Jones polynomial, 3-braids, and the Alexander polynomial We will first derive an expression of the Alexander polynomial of a closed 3-braid in terms of the Jones polynomial. Let Bn be the n-string braid group. The Burau representation of Bn is a map ψ from Bn to n − 1 × n − 1 matrices with entries in Z[t, t−1 ]. ψ : Bn → GL(n − 1, Z[t, t−1 ]). For n = 3, ψ is defined explicitly on the generators σ1 , σ2 (see Figure 5.1) as    −t 1  ψ(σ1−1 ) =  , 0 1 134    1 0  ψ(σ2−1 ) =  . t −t 1 n i Figure 5.1: The generator σi . Note that this convention is the opposite of that of [14]. Let a be an element of B3 , a ˆ be the closed braid, and ea be the exponent sum of a. Note that when a ˆ is a knot, 2 ± ea is even and therefore ea is even. The Jones polynomial Jaˆ (t) of a ˆ can be written in terms of ψ [68]: √ Jaˆ (t) = (− t)−ea (t + t−1 + trace ψ(a)) (5.1) The sign change on ea is due to the difference in convention as indicated in Figure 5.1. When n = 3, the Alexander polynomial of a ˆ may be written in terms of the trace of ψ. [14, Eq. (7)] (t−1 + 1 + t) aˆ (t) = (−1)−ea (t−ea /2 − tea /2 trace ψ(a) + tea /2 ) (5.2) Rearranging equations (5.1) and (5.2) above, we have the symmetric Alexander polynomial of a closed 3-braid re-written in terms of the Jones polynomial. (t−1 + 1 + t) aˆ (t) = (−1)−ea (−(−1)ea tea Jaˆ (t) + tea /2+1 + tea /2−1 + t−ea /2 + tea /2 ). (5.3) This expression allows us to compute certain coefficients of the Alexander polynomial 135 from the Jones polynomial. By Birman and Menasco’s solution [15] to the classification of 3-braids, there are finitely many conjugacy classes of B3 , and each 3-braid is isotopic to a representative of a conjugacy class. Schreier’s work [102] puts each representative of a conjugacy class in a normal form. Theorem 5.1.1. (Schreier) Let b ∈ B3 be a braid on three strands, and C be the 3-braid (σ1 σ2 )3 . Then b is conjugate to a braid in exactly one of the following forms: p −q p −q 1. C k σ1 1 σ2 1 · · · σ1 s σ2 s , where k ∈ Z and pi , qi and s are all positive integers, p 2. C k σ1 , for k, p ∈ Z, 3. C k σ1 σ2 , for k ∈ Z, 4. C k σ1 σ2 σ1 , for k ∈ Z, or 5. C k σ1 σ2 σ1 σ2 , for k ∈ Z. It suffices to study the 3-braids among the conjugacy representatives above to determine p which closed 3-braids is an L-space knot. It is straightforward to check that C k σ1 and C k σ1 σ2 σ1 represent links for any k, p ∈ Z. Also, by noting that C ∼ (σ1 σ2 )3 , we get that, for any k ∈ Z, C k σ1 σ2 and C k σ1 σ2 σ1 σ2 represent the (3, 3k + 1) and (3, 3k + 2) torus knots, respectively. Thus we will only need to study class (1) of conjugacy representatives of Theorem 5.1.1. Recall that if a knot K is a L-space knot, then the absolute value of a nonzero coefficient of the Alexander polynomial K (t) is 1, and the nonzero coefficients alternate in sign. Moreover, let g be the maximum degree of K (t) in t, then the coefficients of the term tg−1 is nonzero and therefore ±1. Since the Alexander polynomial is symmetric, it has the two 136 possible forms given below for an L-space knot. tg − tg−1 + · · · + terms in-between − t−(g−1) + t−g or −tg + tg−1 + · · · + terms in-between + t−(g−1) − t−g . Either way, when we take the product −1 K (t) · (t + 1 + t). The result is a symmetric polynomial with coefficients in {−1, 0, 1}, which do not necessarily alternate in sign, and the second coefficient and the second-to-last coefficient are zero. The conjugacy representatives of class (1) in Theorem 5.1.1 are called generic 3-braids. That is, b ∈ B3 is generic if it has the following form. p −q p −q b = C k σ1 1 σ2 1 · · · σ1 s σ2 s , where pi , qi , k ∈ Z, with pi , qi > 0, and C = (σ1 σ2 σ1 )2 = (σ1 σ2 )3 by the braid relations p −q p −q σ1 σ2 σ1 = σ2 σ1 σ2 . The braid a = σ1 1 σ2 1 · · · σ1 s σ2 s is called an alternating 3-braid. The first three coefficients and the last three coefficients of the Jones polynomial for this class of 3-braids, as well as the degree, are explicitly calculated in [31]. We assemble below the results we will need. 137 p −q p −q Definition 5.1.2. For an alternating braid a = σ1 1 σ2 1 · · · σ1 s σ2 s , let s p := i=1 s pi , and q := i=1 qi , so the exponent sum ea = p − q. Lemma 5.1.3. [31, Lemma 6.2] Suppose that a link a ˆ is the closure of an alternating 3-braid a: p −q p −q a = σ1 1 σ2 1 · · · σ1 s σ2 s , with pi , qi > 0 and p > 1 and q > 1, then the following holds (a) The highest and lowest powers, M (ˆ a) and m(ˆ a) of Jaˆ (t) in t are M (ˆ a) = q − 3p 3q − p and m(ˆ a) = . 2 2 (b) The first two coefficients α, β from M (ˆ a) , and the last two coefficients β , α in Jaˆ (t) from m(ˆ a) are α = (−1)p , β = (−1)p+1 (s − q ), β = (−1)q+1 (s − p ), α = (−1)q . where p = 1 if p = 2 and 0 if p > 2, and similarly for q . (c) [31, in the proof of Lemma 6.2] Let γ, γ denote the third and the third-to-last coefficient of Jaˆ (t), respectively. We have (−1)p γ = s2 + 3s − #{i : pi = 1} − #{i : qi = 1} − δq=3 , 2 138 and (−1)q γ = s2 + 3s − #{i : pi = 1} − #{i : qi = 1} − δp=3 , 2 where δq=3 is zero if q = 3 and 1 otherwise, and δp=3 is similarly defined. The next result writes the Jones polynomial of a generic 3-braid in terms of the Jones polynomial of an alternating braid. Lemma 5.1.4. [31] If b is a generic braid of the form b = C k a, where a is an alternating 3-braid, and let Jˆb (t) denote the Jones polynomial of ˆb, then √ Jˆb (t) = t−6k Jaˆ (t) + (− t)−ea (t + t−1 )(t−3k − t−6k ). If K is a knot which is the closure of a generic 3-braid b = C k a, then by equation (5.3), we have −1 K (t) · (t + 1 + t) = (−1)−eb (−(−1)eb teb Jˆb (t) + teb /2+1 + teb /2−1 + t−eb /2 + teb /2 ). 139 Putting this together with Lemma 5.1.4 and noting that eb must be even in order for ˆb to be a knot, we get that the right side is equal to √ = −teb (t−6k Jaˆ (t) + (− t)−ea (t + t−1 )(t−3k − t−6k )) + teb /2+1 + teb /2−1 + t−eb /2 + teb /2 . Since b = C k a and eb = 6k + ea , we have −1 K (t)(t + 1 + t) √ = −t(6k+ea ) (t−6k Jaˆ (t) + (− t)−ea (t + t−1 )(t−3k − t−6k )) + t(6k+ea )/2+1 + t(6k+ea )/2−1 + t−(6k+ea )/2 + t(6k+ea )/2 = −tea Jaˆ (t) + tea /2−1 + tea /2+1 + t−3k−ea /2 + t3k+ea /2 . 5.2 (5.4) Proof of Theorem 1.1.4 We are now ready to determine which closed 3-braids are L-space knots. Proof. For K = ˆb to be an L-space knot, the right side of equation (5.4) has to be have coefficients in ±1. This immediately restricts s to be less than 4, since s ≥ 4 implies that there will be two coefficients β, β , the second and the penultimate, whose absolute values are greater than or equal to 4 in Jaˆ (t) by (b) of Lemma 5.1.3. Based on (5.4), This will result in at least one coefficient whose absolute value is greater than or equal to 2 in −1 K (t)(t + 1 + t), even after possible cancellation from the terms tea /2−1 , tea /2+1 , t−3k−ea /2 , and t3k+ea /2 . Similarly, if s = 3, then p, q ≥ 3, and |β| and |β | are both greater than or equal to 3, each 140 of which would need to be cancelled out by at least two terms of tea /2−1 , tea /2+1 , t−3k−ea /2 , and t3k+ea /2 . In addition, part (c) of Lemma 5.1.3 gives that the absolute values of the third and third-to-last coefficients γ, γ are greater than or equal to 2, which would also need to be cancelled out in the sum of the right side of (5.4). This is impossible, so s = 3. Now assume that s = 2. If p > 2, then the penultimate coefficient |β | = 2 and the third coefficient |γ| ≥ 2. Since p, q are either both even or both odd, β and γ have opposite signs. One of them is positive, which would not cancel out with any of the terms tea /2−1 , tea /2+1 , t−3k−ea /2 , and t3k+ea /2 . The case is similar for q > 2, so we must have that p = 2 and q = 2. This means that a = σ1 σ2−1 σ1 σ2−1 . The Jones polynomial of this alternating closed braid is Jaˆ(t) = t−2 − t−1 + 1 − t + t2 , for a braid b = C k a, 1 (t−1 + 1 + t) ˆb (t) = −Jaˆ (t) + + t + t−3k + t3k t = −(t−2 − t−1 + 1 − t + t2 ) + 1 + t + t−3k + t3k t = −t−2 + 2t−1 − 1 + 2t − t2 + t−3k + t3k For all k = 0, this shows that the product on the left side of this equation has nonzero coefficients that are not ±1. This rules out the possibility that a closed 3-braid of this form can be an L-space knot. Therefore, we need only to consider the case when s = 1. Assuming that both p, q are greater than 1, the absolute values of the third coefficient γ and the third-to-last coefficient 141 γ of −teα Jαˆ (t) have the form s2 + 3s − #{i : pi = 1} − #{i : qi = 1} − δq=3 2 t(q+p)/2−2 γ and s2 + 3s − #{i : pi = 1} − #{i : qi = 1} − δp=3 2 t−(q+p)/2+2 , γ respectively. If |γ| or |γ | is 2, they need to be canceled out by at least one term out of tea /2−1 , tea /2+1 , t−3k−ea /2 , t3k+ea /2 . If q > 3 and p > 1, then γ = 2, and (q + p)/2 − 2 needs to be equal to (p − q)/2 − 1, (p − q)/2 + 1, −3k − (p − q)/2, or 3k + (p − q)/2. Similarly, we have the constraints on −(q + p)/2 + 2. We examine the resulting equations and rule out values of p and q which lead to contradictions. Setting (q + p)/2 − 2 equal to (p − q)/2 − 1 or (p − q)/2 + 1 gives q = 1 or q = 3. Setting (q + p)/2 − 2 equal to −3k − (p − q)/2 or 3k + (p − q)/2 gives p = −3k + 2 or q = 3k + 2. Similarly, if p > 3 and q > 1, then we must have p = 3, p = 1, q = 3k + 2, or p = −3k + 2. We dismiss the cases where p or q ≤ 3 for now, and suppose that k = 0. We cannot have that p = −3k + 2 and q = 3k + 2 since they are both supposed to be positive. Therefore we suppose that p = −3k + 2 or q = 3k + 2. In the first case, k is 142 negative. In the second case, k is positive. Either way, we end up having, for k < 0, −1 K (t) · (t = (±t− +t + 1 + t) −3k+2+q 2 −3k+2−q −1 2 ∓ t− +t −3k+2+q +1 2 ± 0 ∓ ··· ± 0 ∓ t −3k+2+q −1 2 ±t −3k+2+q 2 ) −3k+2−q +1 2 , or, for k > 0, −1 K (t) · (t = (±t− +t + 1 + t) 3k+2+p 2 ∓ t− −(3k+2)+p −1 2 +t 3k+2+p +1 2 ± 0 ∓ ··· ± 0 ∓ t 3k+2+p −1 2 ±t 3k+2+p 2 ) −(3k+2)+p +1 2 One of the conditions on the product −1 K (t) · (t + 1 + t) is that the second and the penulti- −3k+2+q −3k+2+q +1 −1 2 2 mate coefficients are equal to zero. When k < 0, the terms t− ,t are the second and the penultimate coefficient which is not zero since we assume p, q > 3, this is impossible so this case cannot happen. The same argument applies to rule out the second case when k > 0. When both p, q = 3, we have that the alternating 3-braid a takes the form σ13 σ2−3 . The Alexander polynomial of this alternating 3-braid is a (t) 1 2 = 3 + 2 − − 2t + t2 , t t obtained by multiplying the Alexander polynomial of the trefoil by itself, since this 3-braid is a connected sum of two (right-hand and left-hand) trefoils. It is clear from the Alexander 143 polynomial that this knot cannot be an L-space knot due to the fact that several of its nonzero coefficients are not ±1. Now we consider a generic 3-braid b = C k a with a = σ13 σ2−3 . Since ea = 0, the highest degree and the lowest degree of the Jones polynomial of a ˆ are 3 and −3. By equation (4), 1 (t−1 + 1 + t) ˆb (t) = −Jaˆ (t) + + t + t−3k + t3k , t where 1 1 1 Jaˆ (t) = 3 − 3 + 2 − − t + t2 − t3 . t t t When k = 0, it is clear that the constant term 3 of Jaˆ (t) will not be canceled out by the terms 1t , t, t−3k , or t−3k . Thus none of the closure of braids of the form C k σ13 σ2−3 will be an L-space knot. We may also rule out the case p or q = 2 since this would give a link rather than a knot. Thus, the only generic 3-braids whose closure can be an L-space knot are given below. −q C k σ11 σ2 p C k σ1 σ2−1 for q odd. for p odd. p We now claim that C k σ1 σ2−1 , for p odd and k > 0, represents an L-space knot. Note that: p p (σ1 σ2 σ1 )2k σ1 σ2−1 ∼ (σ2 σ1 )3k σ1 σ2−1 p+1 ∼ (σ2 σ1 )3k−1 σ1 . The latter braid is the twisted torus knot, K(3, 3k − 1; 2, 1), which is known to be an 144 L-space knot [110, Corollary 3.2]. Now if k < 0 then p p (σ1 σ2 σ1 )2k σ1 σ2−1 ∼ (σ1−1 σ2−1 σ1−1 σ1−1 σ2−1 σ1−1 )−k σ1 σ2−1 p ∼ σ2−1 (σ1−1 σ2−1 σ1−1 σ1−1 σ2−1 σ1−1 )−k σ1 p ∼ σ2−1 (σ1−1 σ2−1 σ1−1 σ2−1 σ1−1 σ2−1 )−k σ1 p+1 ∼ σ2−1 (σ1−1 σ2−1 σ1−1 σ2−1 σ1−1 σ2−1 )−k σ1−1 σ1 σ1 p+1 ∼ (σ2−1 σ1−1 )−3k+1 σ1 p+1 ∼ (σ1 σ2 )3k−1 σ1 . Using [110, Corollary 3.2], we get that the closure of the latter braid represents an L-space knot. We should point out that [110, Corollary 3.2], as stated, only holds for positive knots. However, it turns out that every twisted (p, q) torus knot, where the twisting happens between p − 1 strands, admits an L-space surgery from the proof of [110, Theorem 3.1]. A −q similar argument shows that the closure of C k σ11 σ2 , for q odd, also represents an L-space −p−1 knot. Notice that in this case the knot is isotopic to the closure of (σ2 σ1 )1−3k σ1 mirror image admits a positive L-space surgery. 145 . So its BIBLIOGRAPHY 146 BIBLIOGRAPHY [1] Colin Adams and Thomas Kindred. A classification of spanning surfaces for alternating links. Algebr. Geom. Topol., 13(5):2967–3007, 2013. [2] Colin C. Adams. Thrice-punctured spheres in hyperbolic 3-manifolds. Trans. Amer. Math. Soc., 287(2):645–656, 1985. [3] Colin C. Adams. Noncompact Fuchsian and quasi-Fuchsian surfaces in hyperbolic 3–manifolds. Alebr. Geom. Topol., 7:565–582, 2007. [4] Ian Agol. Criteria for virtual fibering. J. Topol., 1(2):269–284, 2008. [5] Ian Agol, Joel Hass, and William Thurston. The computational complexity of knot genus and spanning area. Trans. Amer. Math. Soc., 358(9):3821–3850, 2006. [6] Ian Agol, Peter A. Storm, and William P. Thurston. Lower bounds on volumes of hyperbolic Haken 3-manifolds. J. Amer. Math. Soc., 20(4):1053–1077, 2007. with an appendix by Nathan Dunfield. [7] E. M. Andreev. Convex polyhedra in Lobaˇcevski˘ı spaces. Mat. Sb. (N.S.), 81 (123):445– 478, 1970. [8] E. M. Andreev. Convex polyhedra of finite volume in Lobaˇcevski˘ı space. Mat. Sb. (N.S.), 83 (125):256–260, 1970. [9] Cody Armond. The head and tail conjecture for alternating knots. Algebr. Geom. Topol., 13(5):2809–2826, 2013. [10] Cody Armond and Oliver T. Dasbach. The head and tail of the colored Jones polynomial for adequate knots. arXiv:1310.4537. [11] Cody Armond and Oliver T. Dasbach. Rogers–Ramanujan type identities and the head and tail of the colored Jones polynomial. arXiv:1106.3948. [12] Kenneth L. Baker and Allison H. Moore. Montesinos knots, Hopf plumbings, and L-space surgeries. arXiv:1404.7585. [13] Joshua Batson. Nonorientable four-ball genus can be arbitrarily large. arXiv:1204.1985. 147 [14] Joan Birman. On the jones polynomial of closed 3-braids. Invent. Math., 81(2):287– 294, 1985. [15] Joan S. Birman and Willam W. Menasco. Studying links via closed braids. i. Pacific J. Math., 154(1):17–36, 1992. [16] Gerhard Burde and Heiner Zieschang. Knots, volume 5 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, second edition, 2003. [17] Benjamin A. Burton and Melih Ozlen. Computing the crosscap number of a knot using integer programming and normal surfaces. ACM Trans. Math. Software, 39(1):Art. 4, 18, 2012. [18] Flath Carter, J. Scott, Daniel E., and Saito Masahico. The Classical and Quantum 6j-symbols, volume 43 of Mathematical Notes. Princeton University Press, 1995. [19] Jae Choon Cha and Charles Livingston. Knotinfo: Table of knot invariants. http:// www.indiana.edu/~knotinfo, June 14 2014. [20] Bradd Evans Clark. Crosscaps and knots. Internat. J. Math. Math. Sci., 1(1):0161– 1712, 1978. [21] J. H. Conway. An enumeration of knots and links, and some of their algebraic properties. In Computational Problems in Abstract Algebra (Proc. Conf., Oxford, 1967), pages 329–358. Pergamon, Oxford, 1970. [22] Richard Crowell. Genus of alternating link types. Ann. of Math. (2), 69:258–275, 1959. [23] Oliver T. Dasbach, David Futer, Efstratia Kalfagianni, Xiao-Song Lin, and Neal W. Stoltzfus. The Jones polynomial and graphs on surfaces. Journal of Combinatorial Theory, Series B, 98:384–399, 2008. [24] Oliver T. Dasbach, David Futer, Efstratia Kalfagianni, Xiao-Song Lin, and Neal W. Stoltzfus. Alternating sum formulae for the determinant and other link invariants. J. Knot Theory Ramifications, 19(6):765–782, 2010. [25] Oliver T. Dasbach and Xiao-Song Lin. On the head and the tail of the colored Jones polynomial. Compositio Math., 142(5):1332–1342, 2006. [26] Oliver T. Dasbach and Xiao-Song Lin. A volume-ish theorem for the Jones polynomial of alternating knots. Pacific J. Math., 231(2):279–291, 2007. 148 [27] Nathan Dunfield. Boundary slopes of montesinos knots. http://www.math.uiuc.edu/ ~nmd/montesinos/index.html. [28] Nathan M. Dunfield and Stavros Garoufalidis. Incompressibility criteria for spunnormal surfaces. Trans. Amer. Math. Soc., 364(11):6109–6137, 2012. [29] W. Floyd and U. Oertel. Incompressible surfaces via branched surfaces. Topology, 23(1):117–125, 1984. [30] David Futer and Fran¸cois Gu´eritaud. Angled decompositions of arborescent link complements. Proc. Lond. Math. Soc. (3), 98(2):325–364, 2009. [31] David Futer, Efstratia Kalfagianni, and Jessica Purcell. Cusp areas of farey manifolds and applications to knot theory. Int. Math. Res Not., (23):4434–4497, 2010. [32] David Futer, Efstratia Kalfagianni, and Jessica S. Purcell. Dehn filling, volume, and the Jones polynomial. J. Differential Geom., 78(3):429–464, 2008. [33] David Futer, Efstratia Kalfagianni, and Jessica S. Purcell. Slopes and colored Jones polynomials of adequate knots. Proc. Amer. Math. Soc., 139(5):1889–1896, 2011. [34] David Futer, Efstratia Kalfagianni, and Jessica S. Purcell. Guts of surfaces and the colored Jones polynomial, volume 2069 of Lecture Notes in Mathematics. Springer, Heidelberg, 2013. [35] David Futer, Efstratia Kalfagianni, and Jessica S. Purcell. Jones polynomials, volume, and essential knot surfaces: a survey. In Proceedings of Knots in Poland III, volume 100, pages 51–77. Banach Center Publications, 2014. [36] David Futer, Efstratia Kalfagianni, and Jessica S. Purcell. Quasifuchsian state surfaces. Trans. Amer. Math. Soc., 366:4323–4343, 2014. [37] David Futer and Jessica S. Purcell. Links with no exceptional surgeries. Comment. Math. Helv., 82(3):629–664, 2007. [38] David Gabai. The Murasugi sum is a natural geometric operation. In Low-dimensional topology (San Francisco, Calif., 1981), volume 20 of Contemp. Math., pages 131–143. Amer. Math. Soc., Providence, RI, 1983. [39] David Gabai. Detecting fibred links in S 3 . Comment. Math. Helv., 61(4):519–555, 1986. 149 [40] Stavros Garoufalidis. The degree of a q-holonomic sequence is a quadratic quasipolynomial. Electron. J. Combin., 18(2):Paper 4, 23, 2011. [41] Stavros Garoufalidis. The Jones slopes of a knot. Quantum Topology, 2:43–69, 2011. [42] Stavros Garoufalidis and Thang T. Q. Lˆe. Nahm sums, stability and the colored jones polynomial. Research in Mathematical Sciences, to appear. [43] Stavros Garoufalidis and Thang T. Q. Lˆe. The colored Jones function is q-holonomic. Geom. Topology., (9):1253–1293, 2005. [44] Stavros Garoufalidis and Thang T. Q. Lˆe. Nahm sums, stability and the colored Jones polynomial. Research in Mathematical Sciences, to appear. [45] Stavros Garoufalidis, Sergei Norin, and Thao Vong. Flag algebras and the stable coefficients of the Jones polynomial. arXiv:1309.5867. [46] Stavros Garoufalidis and Roland van der Veen. Quadratic integer programming and the slope conjecture. arXiv:1405.5088. [47] Stavros Garoufalidis and Thao Vong. A stability conjecture for the colored Jones polynomial. arXiv:1310.7143. [48] Paolo Ghiggini. Knot Floer homology detects genus-one fibred knots. American Journal of Mathematics, 130(5):1151–1169, 2008. [49] Fran¸cois Gu´eritaud and David Futer (appendix). On canonical triangulations of oncepunctured torus bundles and two-bridge link complements. Geom. Topol., 10:1239– 1284, 2006. [50] Wolfgang Haken. Theorie der Normalfl¨achen. Acta Math., 105:245–375, 1961. [51] Joel Hass, Jeffrey C. Lagarias, and Nicholas Pippenger. The computational complexity of knot and link problems. J. ACM, 46(2):185–211, 1999. [52] Allen Hatcher. Notes on basic 3-manifold topology. [53] Allen Hatcher. Notes on basic 3-manifold topology. http://www.math.cornell.edu/ ~hatcher/3M/3Mdownloads.html. 150 [54] Allen Hatcher. On the boundary curves of incompressible surfaces. Pacific J. Math., 99(2):373–377, 1982. [55] Allen Hatcher and U. Oertel. 28(4):453–480, 1989. Boundary slopes for Montesinos knots. Topology, [56] Allen Hatcher and William Thurston. Incompressible surfaces in 2-bridge knot complements. Invent. Math., 79(2):225–246, 1985. [57] Matthew Hedden. On knot Floer homology and cabling II. International Mathematics Research Notices, pages 2248–2274, 2009. [58] Matthew Hedden. Notions of positivity and the Oszvath-Szabo concordance invariant. J. Knot Theory Ramifications, 19(5):617–629, 2010. [59] Matthew Hedden. Notions of positivity and the Ozsv´ath-Szab´o concordance invariant. J. Knot Theory Ramifications, 19(5):617–629, 2010. [60] Matthew Hedden and Liam Watson. On the geography and botany of knot Floer homology. preprint, page arXiv:1404.6913, 2014. [61] Mikami Hirasawa and Masakazu Teragaito. Crosscap numbers of 2-bridge knots. Topology, 45(3):513–530, 2006. [62] Jennifer Hom. A note on cabling and L-space surgeries. Algebraic & Geometric topology, (1):219–223, 2011. [63] Jennifer Hom, Faramarz Vafaee, and Tye Lidman. Berge-Gabai knots and L-space satellite operations. Algebraic & Geometric topology, to appear. [64] Kazuhiro Ichihara and Shigeru Mizushima. Crosscap numbers of pretzel knots. Topology Appl., 157(1):193–201, 2010. [65] Kazuhiro Ichihara, Masahiro Ohtouge, and Masakazu Teragaito. Boundary slopes of non-orientable Seifert surfaces for knots. Topology Appl., 122(3):467–478, 2002. [66] William Jaco and J. Hyam Rubinstein. PL equivariant surgery and invariant decompositions of 3-manifolds. Adv. in Math., 73(2):149–191, 1989. [67] Vaughan F. R. Jones. A polynomial invariant for knots via von Neumann algebras. Bull. Amer. Math. Soc. (N.S.), 12(1):103–111, 1985. 151 [68] Vaughan F. R. Jones. Hecke algebra representations of braid groups and link polynomials. Ann. of Math. (2), 126(2):335–388, 1987. [69] Efstratia Kalfagianni and Christine Ruey Shan Lee. On the degree of the colored Jones polynomial. Acta Mathematica Vietnamica (Proceedings of Quantum Topology and Hyperbolic Geometry in Nha Trang, May 2013), to appear. [70] Efstratia Kalfagianni and Anh T. Tran. Knot cabling and the degree of the colored jones polynomial. http://arxiv.org/pdf/1501.01574v2.pdf. [71] Efstratia Kalfagianni and Anh T. Tran. Knot cabling and the degree of the colored jones polynomial ii. [72] Louis H. Kauffman. State models and the Jones polynomial. Topology, 26(3):395–407, 1987. [73] Mikhail Khovanov. A categorification of the Jones polynomial. Duke Mathematical Journal, 101(3):359–426, 2000. [74] David Krcatovich. The reduced knot Floer complex. arXiv:1310.7624. [75] Marc Lackenby. Word hyperbolic dehn surgery. Invent. Math., 140(2):243–282, 2000. [76] Marc Lackenby. The volume of hyperbolic alternating link complements. Proc. London Math. Soc. (3), 88(1):204–224, 2004. With an appendix by Ian Agol and Dylan Thurston. [77] Thang T. Q. Lˆe. The colored Jones polynomial and AJ conjecture. http://people. math.gatech.edu/%7Eletu/Papers/Lectures_Luminy_2014_new.pdf. [78] W. B. Raymond Lickorish. An Introduction to Knot Theory. Springer-Verlag New York, Inc., 1997. [79] W. B. Raymond Lickorish and Morwen B. Thistlethwaite. Some links with nontrivial polynomials and their crossing-numbers. Comment. Math. Helv., 63(4):527–539, 1988. [80] Tye Lidman and Allison H. Moore. arXiv:1306.6707. Pretzel knots with l-space surgeries. [81] P. M. G. Manch´on. Extreme coefficients of Jones polynomials and graph theory. J. Knot Theory Ramifications, 13(2):277–295, 2004. 152 [82] G. Masbaum and P. Vogel. 3-valent graphs and the Kauffman bracket. Pacific J. Math., 164(2):361–381, 1994. [83] William W. Menasco. Closed incompressible surfaces in alternating knot and link complements. Topology, 23(1):37–44, 1984. [84] William W. Menasco and Morwen B. Thistlethwaite. Surfaces with boundary in alternating knot exteriors. J. Reine Angew. Math., 426:47–65, 1992. [85] William W. Menasco and Morwen B. Thistlethwaite. The classification of alternating links. Ann. of Math. (2), 138(1):113–171, 1993. [86] Kimihiko Motegi. L-space surgery and twisting operation. arXiv:1405.6487. [87] Hitoshi Murakami. An introduction to the volume conjecture. In Interactions between hyperbolic geometry, quantum topology and number theory, volume 541 of Contemp. Math., pages 1–40. Amer. Math. Soc., Providence, RI, 2011. [88] Hitoshi Murakami and Jun Murakami. The colored Jones polynomials and the simplicial volume of a knot. Acta Math., 186(1):85–104, 2001. [89] Hitoshi Murakami and Akira Yasuhara. Crosscap number of a knot. Pacific J. Math., 171(1):261–273, 1995. [90] Kunio Murasugi. On the Alexander polynomial of the alternating knot. Osaka Math. J., 10:181–189; errata, 11 (1959), 95, 1958. [91] Kunio Murasugi. Jones polynomials and classical conjectures in knot theory. Topology, 26(2):187–194, 1987. [92] Yi Ni. Knot Floer homology detects fibred knots. Invent. Math., 170(3):577–608, 2007. [93] Makoto Ozawa. Essential state surfaces for knots and links. J. Aust. Math. Soc., 91(3):391–404, 2011. [94] Peter Ozsv´ath and Zolt´an Szab´o. Holomorphic disks and genus bounds. Geom. Topol., 8:311–334, 2004. [95] Peter Ozsv´ath and Zolt´an Szab´o. On knot Floer homology and lens space surgeries. Topology, 44(6):1281–1300, 2005. 153 [96] Peter Ozsvath and Zoltan Szabo. On the heegaard floer homology of branched doublecovers. Advances in Mathematics, 194(1):1–33, 2005. [97] Peter Ozsv´ath and Zolt´an Szab´o. Link Floer homology and the Thurston norm. J. Amer. Math. Soc., 21(3):671–709, 2008. [98] Jessica S. Purcell. An introduction to fully augmented links. In Interactions between hyperbolic geometry, quantum topology and number theory, volume 541 of Contemp. Math., pages 205–220. Amer. Math. Soc., Providence, RI, 2011. [99] N. Yu. Reshetikhin and V. G. Turaev. Ribbon graphs and their invariants derived from quantum groups. Comm. Math. Phys., 127(1):1–26, 1990. [100] Nikolai Reshetikhin and Vladimir G. Turaev. Invariants of 3-manifolds via link polynomials and quantum groups. Invent. Math., 103(3):547–597, 1991. [101] Lev Rozansky. Khovanov homology of a unicolored B-adequate link has a tail. arXiv:1203.5741. [102] Otto Schreier. U¨ ber die gruppen aabb = 1. Abh. Math. Sem. Univ. Hamburg, (3):167– 169, 1924. [103] Hermann Schubert. Knoten mit zwei brucken. Math. Zeitschr, 65:133–170, 1956. [104] Alexander Stoimenow. Non-triviality of the jones polynomial and the crossing numbers of amphicheiral knots. [105] Alexander Stoimenow. Coefficients and non-triviality of the Jones polynomial. Journal f¨ ur die Reine und Angewandte Mathematik, 657:1–55, 2011. [106] Alexander Stoimenow. On the crossing number of semi-adequate links. Forum Math., pages DOI:10.1515/forum–2011–0121, in press. [107] Masakazu Teragaito. Crosscap numbers of torus knots. Topology Appl., 138(1-3):219– 238, 2004. [108] Morwen B. Thistlethwaite. On the Kauffman polynomial of an adequate link. Invent. Math., 93(2):285–296, 1988. 154 [109] Vladimir G. Turaev. Quantum invariants of knots and 3-manifolds, volume 18 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, revised edition, 2010. [110] Faramarz Vafaee. On the knot Floer homology of twisted torus knots. Int. Math. Res. Not. IMRN, page doi: 10.1093/imrn/rnu130, 2014. 155