A GEOMETRIC PROOF OF THE FLYPING THEOREM

THOMAS KINDRED

Abstract. In 1898, Tait asserted several properties of alter- nating diagrams. These assertions came to be known as Tait’s conjectures and remained open through the discovery of the in 1985. The new polynomial invari- ants soon led to proofs of all of Tait’s conjectures, culminating in 1993 with Menasco–Thistlethwaite’s proof of Tait’s flyping conjecture. In 2017, Greene gave the first geometric proof of part of Tait’s conjectures, in the same paper where he answered a long- standing question of Fox by characterizing alternating links ge- ometrically; Howie independently answered Fox’s question with a related characterization. We use Greene and Howie’s charac- terizations, Menasco’s crossing ball structures, and re-plumbing moves to give the first entirely geometric proof of Menasco– Thistlethwaite’s Flyping Theorem.

1. Introduction P.G. Tait asserted in 1898 that any two reduced alternating dia- grams of a prime non-split in S3 are related by a sequence of flype moves (see Figure 1) [11]. Tait’s flyping conjecture remained open until 1993, when Menasco–Thistlethwaite gave its first proof [8, 9], which they described as follows: The proof of the Main Theorem stems from an anal-

arXiv:2008.06490v1 [math.GT] 14 Aug 2020 ysis of the [chessboard surfaces] of a link diagram, in which we use geometric techniques [introduced in [7]]... and properties of the Jones and Kauffman poly- nomials.... Perhaps the most striking use of polyno- mials is... where we “detect a flype” by using the fact that if just one crossing is switched in a reduced alter- nating diagram of n crossings, and if the resulting link also admits an alternating diagram, then the crossing number of that link is at most n − 2. Thus, although the proof of the Main Theorem has a strong geomet- ric flavor, it is not entirely geometric; the question 1

2 THOMAS KINDRED

T 2 T1 T2 T1

Figure 1. A flype along a curve γ. Two link dia- grams are flype-related if there is a sequence of flype moves taking one diagram to the other.

remains open as to whether there exist purely geo- metric proofs of this and other results that have been obtained with the help of new polynomial invariants. We answer part of this question by giving the first entirely geo- metric proof of the Flyping Theorem. (Greene answered part of the question in [3], by providing the first geometric proof of part of Tait’s conjectures. Other parts of the question remain open; see Problems 2.6–2.8.) Flyping Theorem. Any two reduced alternating diagrams of the same prime, nonsplit link are flype-related. Like Menasco-Thistlethwaite’s proof, ours stems from an analy- sis of chessboard surfaces and uses the geometric techniques intro- duced in [7]. Recent insights of Greene and Howie regarding these chessboards also play a central role in our analysis [3, 4]. (Those in- sights answered another longstanding question, this one from Ralph Fox: “What [geometrically] is an [or link]?”) Per- haps the most striking difference between our proof and Menasco- Thistlethwaite’s is that we “detect flypes” via re-plumbing moves: Theorem 4.11. Let D and D0 be reduced alternating diagrams of a prime nonsplit link L with respective chessboards B,W and B0,W 0, where B and B0 are positive-definite. Then D and D0 are flype-related if and only if B and B0 are related by isotopy and re-plumbing moves, as are W and W 0.1 Section 2 discusses chessboards and definite surfaces, and §3 ad- dresses generalized plumbing and re-plumbing moves.2 Figure 13 shows the type of re-plumbing move associated with flyping. The- orem 4.11 is the first of two intermediate results which lead to the

1The proof of Theorem 4.11 shows that a reduced alternating diagram D00 of L with chessboards isotopic to B0 and W is related to D (resp. D0) by a sequence of flypes each of which corresponds to an isotopy of W (resp. B0). 2Generalized plumbing is also called Murasugi sum. A GEOMETRIC PROOF OF THE FLYPING THEOREM 3 proof of the Flyping Theorem. With the same setup as Theorem 4.11, the second of these results is:3 Theorem 5.26. Any essential positive-definite surface spanning L is related to B by isotopy and re-plumbing moves.4 Together with background from [3], Theorems 4.11 and 5.26 im- mediately imply the Flyping Theorem. Section 4 sets up crossing ball structures and uses them to prove Theorem 4.11. Section 5 uses these structures to prove Theorem 5.26 and the Flyping Theorem. Several of the arguments in Sections 4 and 5 are adapted from [1, 5].

2. Definite surfaces 2.1. Background. Let L be a link in S3 with a closed regular neigh- 5 borhood νL and projection πL : νL → L. One can define spanning surfaces for L in two ways; in both definitions, F is compact, but not necessarily orientable, and each component of F has nonempty boundary.6 First, a spanning surface for L is an embedded surface F ⊂ S3 with ∂F = L. Alternatively, a spanning surface is a prop- ◦ erly embedded surface F in the link exterior S3 \ νL such that ∂F intersects each meridian on ∂νL transversally in one point.7 We use the latter definition. The rank β1(F ) of the first homology group of a spanning sur- face F counts the number of “holes” in F . When F is connected, β1(F ) = 1 − χ(F ) counts the number of cuts — along disjoint, prop- erly embedded arcs — required to reduce F to a disk. Thus: Observation 2.1. If α is a properly embedded arc in a spanning 0 ◦ 0 surface F and F = F \ να is connected, then β1(F ) = β1(F ) + 1. Given any diagram D of L, one may a color the complementary regions of D in the projection sphere S2 black and white in chess- board fashion.8 One may then construct spanning surfaces B and W for L such that B projects into the black regions, W projects into the white, and B and W intersect in vertical arcs which project to the the crossings of D. Call the surfaces B and W the chessboards from D. Figure 2 illustrates chessboards near a crossing. Given any spanning surface F for a link L ⊂ S3, Gordon-Litherland construct a symmetric, bilinear pairing h·, ·i : H1(F ) × H1(F ) → Z 3In this setup, both B and W are essential (see §2.3 for the definition), since D is reduced alternating and L is nonsplit. 4Likewise for negative-definite surfaces and W . 5Convention: Given X ⊂ S3, νX denotes a closed regular neighborhood of X and is taken in S3 unless stated otherwise. 6When L is nonsplit and alternating, F is connected, by Corollary 5.2 of [1]. 7 −1 Definition: A meridian on ∂νL is a circle πL (x) ∩ ∂νL for a point x ∈ L. 8That is, so that regions of the same color meet only at crossing points. 4 THOMAS KINDRED

Figure 2. Chessboards near a vertical arc at a crossing. as follows [2]. Let νF be a neighborhood of F in the link exterior S3 \\νL with projection p : νF → F , such that p−1(∂F ) = νF ∩∂νL; denote Fe = ∂νF \\∂νL.9 Given any oriented simple closed curve −1 γ ⊂ F , denote γe = ∂(p (γ)), and orient γe following γ. Let τ : H1(F ) → H1(Fe) be the transfer map characterized by 10 p∗ ◦ τ = 2 · 1. The Gordon-Litherland pairing

h·, ·i : H1(F ) × H1(F ) → Z is the symmetric, bilinear mapping given by the ha, bi = lk(a, τ(b)). The matrix representing h·, ·i, with respect to an ordered basis B = m×m (a1, . . . , am) for H1(F ), is the Goeritz matrix G = (xij) ∈ Z 11 given by xij = hai, aji. Any projective homology class g ∈ H1(F )/± has a well-defined self-pairing hg, gi. When F is connected and g is primitive, there 1 is a simple closed curve γ ⊂ F representing g, and 2 hg, gi equals the framing of γ in F . A spanning surface F is positive-definite if hα, αi > 0 for all nonzero α ∈ H1(F ). The spectral theorem implies that G has β1(F ) real eigenvalues, counted with multiplicity; thus, F is positive-definite iff all eigenval- ues of G are positive. When F is connected, an equivalent condition,

9For compact X,Y ⊂ S3, X \\Y denotes the metric closure of X \ Y . In case Z ∩ ν(x) is connected or empty for each component Z of X \ Y and x ∈ X ∩ Y , each component of X \\Y is the closure in S3 of a component of X \ Y , hence embeds naturally in S3, although X \\Y may not. Also see Note 16. 10 In other words, given any primitive g ∈ H1(F ), choose an oriented simple closed curve γ ⊂ int(F ) representing g, and construct γe as above; then, τ(g) = [γe]. 11 Pm Pm That is, any y = i=1 yiai and z = i=1 ziai satisfy    T hy, zi = y1 ··· ym G z1 ··· zm . A GEOMETRIC PROOF OF THE FLYPING THEOREM 5 more geometric in flavor, is that F is positive-definite iff, for each simple closed curve γ ⊂ F , either: • The framing of γ in F is positive, or • γ bounds an orientable subsurface of F . Negative-definite surfaces are defined analogously. Greene proves: Theorem 1.1 of [3]. If B and W are positive- and negative-definite spanning surfaces for a link L in a homology Z/2Z sphere with irre- ducible complement, then L is an alternating link in S3, and it has an alternating diagram whose chessboards are isotopic to B and W . Moreover, this diagram is reduced if and only if neither B nor W has a homology class with self-pairing ±1.12 Convention 2.2. Isotopies of properly embedded surfaces and arcs are always taken rel boundary.13 Two properly embedded surfaces or arcs are parallel if they have the same boundary and are related by an isotopy which fixes this boundary. The converse of the first sentence of the theorem is also true: Proposition 4.1 of [3]. A connected link diagram is alternating if and only if its chessboards are definite surfaces of opposite signs. Convention 2.3. If D is a connected alternating link diagram, then its chessboards B and W are labeled such that B is positive-definite and W is negative-definite. Likewise for chessboards B0 and W 0 0 (resp. Bi and Wi) from such a diagram D (resp. Di). The euler number e(F ) of a spanning surface F is the algebraic self- intersection number of the properly embedded surface in the 4-ball obtained by perturbing F . Alternatively, −e(F ) can be computed by summing the component-wise boundary slopes of F .14 For brevity, we call −e(F ) the slope of F and denote −e(F ) = s(F ). Proposition 2.4. If D and D0 are reduced alternating diagrams of the same prime nonsplit link, and if B,W and B0,W 0 are their re- spective chessboards, then s(B) = s(B0) and s(W ) = s(W 0). Greene proves Proposition 2.4 as a step in his proof of Theorem 1.2. Here, we prove, alternatively, that the proposition can be seen as a consequence of the theorem. Proof. In general, the slopes of chessboards can be computed from the signs of the crossings in the diagram. In the alternating case, the

12A diagram D is reduced if each crossing abuts four distinct regions of S2 \D. 13For example, an isotopy of a spanning surface F ⊂ S3 \\νL is a homotopy h : F × I → S3 \\νL for which h(F × {t}) is a spanning surface at each t ∈ I. 14 Namely, if L1,...,Lm are the components of ∂F , and `1, . . . , `m are their 1 Pm projective homology classes in H1(F )/±, then −e(F ) = 2 i=1h`i, `ii. 6 THOMAS KINDRED computation is particularly nice. Namely:

s(B) = 2 · and s(W ) = −2 · , D D and likewise for B0 and W 0. Since c(D) = c(D0) andd w(D) = w(D0) by Theorem 1.2 of [3], we have: s(B) − s(W )

= + = c(D) (1) 2 D D s(B0) − s(W 0) = c(D0) = + = D0 D0 2 and s(B) + s(W )

= − = w(D) (2) 2 D D s(B0) + s(W 0) = w(D0) = − = . D0 D0 2 0 0 Combining (1) and (2) gives s(B) = s(B ) and s(W ) = s(W ).  Greene uses Theorem 1.1 of [3] to give a geometric proof of part of Tait’s conjectures: Theorem 1.2 of [3]. Any two reduced alternating diagrams of the same link have the same crossing number and . Remark 2.5. This does not imply, a priori, that a reduced alter- nating diagram realizes the underlying link’s crossing number. All existing proofs of this fact [6, 10, 12, 15] use the Jones polynomial. The following problem remains open: Problem 2.6. Give an entirely geometric proof that any reduced al- ternating link diagram realizes the underlying link’s crossing number. More generally, Corollary 3.4 of [13] states that any adequate link diagram realizes the crossing number of the underlying link. Problem 2.7. Prove Corollary 3.4 of [13] geometrically. There is a closely related open problem for tangles. Theorem 3.1 of [14] states that any reduced alternating diagram realizes the underlying tangle’s crossing number. See [14] for the definitions, and for a generalized formulation of the theorem. Problem 2.8. Give a geometric proof of Theorem 3.1 of [14]. 2.2. Operations on definite surfaces. Definite surfaces behave well under boundary-connect sum: Proposition 2.9. If F is a boundary-connect sum of positive-definite surfaces F1 and F2, then F is positve-definite. A GEOMETRIC PROOF OF THE FLYPING THEOREM 7

Proof. Let Gi be a Goeritz matrix for Fi for i = 1, 2. Then each Gi has β1(Fi) positive eigenvalues (counted with multiplicity). Further, G 0  G = 1 0 G2 is a Goeritz matrix for F with β1(F ) positive eigenvalues.  Next, consider subsurfaces of definite surfaces. Greene proves: Lemma 3.3 of [3]. If F is a definite surface and F 0 ⊂ F is a compact subsurface with connected boundary, then F 0 is definite. Here is a related fact: Proposition 2.10. If F 0 is a compact subsurface of a definite sur- face F , and every component of F \ F 0 intersects ∂F , then every component of F 0 is definite. 0 0 0 Proof. Let F0 be a component of F , let j : F0 → F denote inclusion, 0 and let g ∈ H1(F0) be primitive with j∗(g) = 0 ∈ H1(F ). Choose a 0 00 simple closed curve γ ⊂ F0 representing g. Then γ = ∂F for some 00 0 00 0 orientable F ⊂ F . The fact that γ ⊂ F0 implies that F ⊂ F , or else F 00 would intersect, and therefore contain, some component of 0 00 F \ F , implying contrary to assumption that ∂F 6= γ. Ergo, j∗ is 0 injective, so F is definite.  In particular, Proposition 2.10 immediately implies: F Lemma 2.11. If F is a definite surface and α = i∈I αi is a system of disjoint, properly embedded arcs in F , then every component of ◦ F \ να is definite. Note also: Proposition 2.12. Let F be a spanning surface for a prime, nonsplit link L ⊂ S3 and α a properly embedded arc in F . If α is not ∂-parallel ◦ in S3 \\νL, then the link L0 = ∂(F \ να) is nonsplit. Proof. Assume for contradiction that L0 is split. Then there is a sphere Q ⊂ S3 \ L0 which intersects F along α, and perhaps some simple closed curves. The fact that L is prime implies, contrary to 3 assumption, that α is ∂-parallel in S \\νL.  Next, consider the operation of adding (half) twists, illustrated in Figure 3. It works like this. Let F be a spanning surface for a link L, α ⊂ F a properly embedded arc which is not ∂-parallel in F , m a nonzero integer. Let A be an unknotted annulus or mobius band whose core circle has framing m, and let α0 be a properly embedded ◦ arc in A such that A \ να0 is a disk. Construct a boundary connect sum F \A in such a way that α and α0 are glued together to form an 8 THOMAS KINDRED

Figure 3. To add m half twists to a spanning surface F (bottom left) along an arc α ⊂ F , take the band A with cocore α0 (top left) spanning the torus link T2,m, construct F \A (center) so as to glue α and α0 into an arc α00, and cut F \A along α00.

◦ arc α00. Depending on the sign of m, the surface F 0 = (F \A) \ να00 is said to be obtained from F by adding (positive or negative) twists along α. Proposition 2.13. If F 0 is obtained by adding positive twists to a positive-definite surface F , then F 0 is positive-definite.15 Indeed, if G is a positive-definite symmetric matrix and G0 is ob- tained by increasing a diagonal entry of G, then G0 is also positive- definite. Alternatively, here is a geometric proof:

Proof. Let A be an unknotted annulus or mobius band with m half- twists for some m > 0. Then A is also positive-definite, as are F \A 0 and F , by Proposition 2.9 and Lemma 2.11. 

2.3. Essentiality. A spanning surface F is (geometrically) incom- pressible if any simple closed curve in F that bounds a disk in S3 \\(F ∪ νL) also bounds a disk in F .16 A spanning surface F is ∂-incompressible if any properly embedded arc in F that is ∂-parallel in S3 \\(F ∪ νL) is also ∂-parallel in F . If F is incompressible and ∂-incompressible, then F is essential. This geometric notion of essen- tiality is weaker than the algebraic notion of π1-essentiality, which

15Likewise for adding negative twists to a negative-definite surface. 16 Continuing Note 9, here is a general construction for X \\Y . Let {(Uα, φα)} be a maximal atlas for X. About each x ∈ X, choose a chart (Ux, φx) such that Ux is a regular open neighborhood of x with closure Ux. Construct Ux \\Y as in Note 9; denote the components of Ux ∩ (Ux \\Y ) by Uα, α ∈ Ix; denote each natural embedding fα : Uα → Ux. Then X \\Y has the following atlas: [ {(Uα, φx ◦ fα)}α∈Ix x∈X

3 Gluing all the maps fα yields a natural map f : X \\Y → X ⊂ S . A GEOMETRIC PROOF OF THE FLYPING THEOREM 9 holds F to be essential if F is not a mobius band spanning the un- knot, and if inclusion F,→ S3 \\νL induces an injective map on fundamental groups. Recall from Theorem 1.1 of [3] that a connected alternating di- agram is reduced iff neither chessboard has a homology class with self-pairing ±1. An equivalent condition is that both chessboards are essential. Proposition 2.14. If an essential surface F contains a properly embedded arc β which is parallel in S3 \\(F ∪ νL) to a properly embedded arc α ⊂ ∂νL \\∂F , then α is parallel in ∂νL to ∂F . Proof. Since F is essential and β is parallel in the link exterior to α ⊂ ∂νL, β is also parallel in F to an arc α0 ⊂ ∂F . Thus, α and α0 are both parallel in S3 \\(F ∪ νL) to β, and so α is parallel through a disk X ⊂ S3 \\(F ∪ νL) to α0 ⊂ ∂F . Since L is nontrivial and nonsplit, X is parallel in S3 \\(F ∪ νL) to a disk X0 ⊂ ∂νL, through 0 which α is parallel to α ⊂ ∂F . 

2.4. Signs and slopes. Let F and F 0 be spanning surfaces for a link 3 0 0 L ⊂ S with F t F and in particular ∂F t ∂F on ∂νL. Orient L arbitrarily, and orient ∂F and ∂F 0 so that each is homologous in νL 0 to L. Then the algebraic intersection number i(∂F, ∂F )∂νL equals the (net) slope difference s(F ) − s(F 0) (see §2.1). Independently from Greene, Howie characterized alternating and links as follows: Theorem 3.1 of [4]. Let L be a nontrivial knot in S3 with exterior X. Then L is alternating if and only if there exist a pair of connected spanning surfaces Σ+, Σ− in X which satisfy 1 (3) χ(Σ ) + χ(Σ ) + i(∂Σ , ∂Σ ) = 2. + − 2 + − ∂νL

Theorem 3.2 of [4]. Let L = L1 t · · · t Lm be a nontrivial nonsplit link in S3 with exterior X which has a marked meridian on each boundary component. Then L has an alternating projection onto S2 if and only if there exist a pair of connected spanning surfaces Σ+, Σ− in X which satisfy (3) and m X i(∂Σ+, ∂Σ−)∂νL = i(∂Σ+, ∂Σ−)∂νLj . j=1

Howie shows that the surfaces Σ± can always be isotoped to in- tersect only in arcs, and that when they do Σ± can be seen as chess- boards from an alternating diagram of L. By combining Howie’s characterization with Greene’s, we have: 10 THOMAS KINDRED

Lemma 2.15. If F+ and F−, respectively, are positive- and negative- definite spanning surfaces for the same nonsplit link L ⊂ S3, then

s(F+) − s(F−) = 2(β1(F+) + β1(F−)).

If F+ and F− are essential, then s(F+) − s(F−) = 2n, where n is the number of crossings in any reduced alternating diagram of L.

Proof. If neither F+ nor F− has a homology class with self-pairing ±1, then any diagram with F+ and F− as chessboards is reduced by Theorem 1.1 of [3]; thus, s(F+) − s(F−) = 2n. Also, in that case, β1(F+) + β1(F−) = n, so the result follows. This extends to the general case, in which F+ has some number of positive crosscaps attached, and F− some negative ones, by noting that attaching a positive or negative crosscap increases β1 by one and changes slope 17 by ±2, respectively.  In the following way, one can compute the slope difference s(F ) − s(F 0) by counting, with signs, the arcs of F ∩ F 0. Given an arc α of F ∩ F 0, let ν∂α be a regular neighborhood of the endpoints ∂α in ∂νL. Then ∂F ∩ ν∂α and ∂F 0 ∩ ν∂α0 each consist of two properly embedded arcs, one in each disk of ν∂α. The two arcs in each disk intersect transversally in a single point, giving 0 i(∂F, ∂F )ν∂α ∈ {0, ±2}. 0 Partition the arcs of F ∩ F as A2 ∪ A0 ∪ A−2, such that each  0 0 Aj = arcs α of F ∩ F : i(∂F, ∂F )ν∂α = j . Then: 0 (4) s(F ) − s(F ) = 2|A2| − 2|A−2|. Combining (4) with several earlier results gives: Lemma 2.16. If F and W respectively are positive- and negative- definite surfaces spanning a prime nonsplit link L ⊂ S3 and α is an arc of F ∩ W which is not ∂-parallel in the link exterior, then i(∂F, ∂W )ν∂α = 2.

Proof. Assume for contradiction that i(∂F, ∂W )ν∂α 6= 2. 0 ◦ First, consider the case i(∂F, ∂W )ν∂α = −2. Let F = F \ να, and let W 0 be the surface obtained by adding one (negative) half- twist to W along α. Then F 0 and W 0 span the same link L0, with 0 0 β1(F ) = β1(F ) − 1 and β1(W ) = β1(W ). See Figure 4.

17A similar argument uses Lemma 2.15 to show that any positive-definite surfaces F and F 0 spanning the same nonsplit alternating link L ⊂ S3 satisfy 0 0 0 s(F ) − s(F ) = 2(β1(F ) − β1(F )), and any negative-definite surfaces F and F 0 0 spanning L satisfy s(F ) − s(F ) = 2(β1(F ) − β1(F )). A GEOMETRIC PROOF OF THE FLYPING THEOREM 11

−→

Figure 4. A positive-definite surface F cannot in- tersect a negative-definite surface W along a non-∂- parallel arc α with i(∂F, ∂W )ν∂α = −2.

Lemma 2.11 implies that F 0 is positive-definite, and Proposition 2.13 implies that W 0 is negative-definite. Hence, L0 is alternating, by Theorem 1.1 of [3]. Further, L0 is nonsplit by Proposition 2.12. Therefore, by Corollary 5.2 of [1], F 0 and W 0 are both connected. Thus, by Observation 2.1 and (4): s(F 0) − s(W 0) = s(F ) − s(W ) + 2

(5) = 2(β1(F ) + β1(W )) + 2 0 0 > 2(β1(F ) + β1(W )). This contradicts Lemma 2.15. Now, consider the case i(∂F, ∂W )ν∂α = 0. The argument here is identical to the first case, except that one constructs W 0 by cutting ◦ 0 out να, rather than adding twists, giving β1(F ) = β1(F ) − 1 and 0 β1(W ) = β1(W ) − 1 s(F 0) − s(W 0) = s(F ) − s(W )

(6) = 2(β1(F ) + β1(W )) 0 0 > 2(β1(F ) + β1(W )).

Again, this contradicts Lemma 2.15. 

3. Generalized plumbing Let F be a spanning surface for a nonsplit link L (not necessarily alternating). A plumbing cap for F is a properly embedded disk V ⊂ S3 \\(F ∪ νL) with the following properties: • The natural map f : S3 \\(F ∪ νL) → S3 restricts to an embedding f : V → S3, and f(∂V ) bounds a disk Ub ⊂ F ∪νL. 3 • Denoting the 3-balls of S \\(Ub ∪ V ) by Y1,Y2, neither sub- surface Fi = F ∩ Yi is a disk. 12 THOMAS KINDRED

= along

Figure 5. A plumbing cap and its shadow for a span- ning surface, and the associated de-plumbing.

The disk U = Ub ∩ F is the shadow of V . If the first property holds but the second fails, we call V a fake plumbing cap for F ; we still call U the shadow of V . The decomposition F = F1 ∪ F2 is a plumbing decomposition or de-plumbing of F along U and V , denoted F = F1 ∗ F2. See Figure 5. The reverse operation, in which one glues F1 and F2 along U to produce F , is called generalized plumbing or Murasugi sum. If V is a plumbing cap for F with shadow U, then one can construct another spanning surface F 0 = (F \ U) ∪ V ; we call the operation of changing F to F 0 re-plumbing F along U and V . Call the analogous operation along a fake plumbing cap a fake re-plumbing. This is an isotopy move. Two spanning surfaces are plumb-related if there is sequence of re-plumbing and isotopy moves taking one surface to the other. Proposition 3.1. If V is a plumbing cap or fake plumbing cap for a spanning surface F , and U is the shadow of V , then F is plumb- related to (F \\U) ∪ V . Proof. If V is a fake plumbing cap, then F is isotopic to (F \\U)∪V . Otherwise, a re-plumbing move takes F to (F \\U) ∪ V .  Definition 3.2. A plumbing cap V is acceptable if both: • No arc of ∂V ∩ ∂νL is parallel in ∂νL to ∂F . • No arc of ∂V ∩ F is parallel in F \\U to ∂F . Any, possibly fake, plumbing cap V for F can be adjusted near any arcs described in Definition 3.2 so as to remove those arcs one at a time. Doing so for a plumbing cap eventually yields an acceptable plumbing cap V 0; in the fake case, an F -parallel disk. Ergo: Observation 3.3. No fake plumbing cap satisfies both properties from Definition 3.2. That is, there is no such thing as an acceptable fake plumbing cap. With V 0 as above: Observation 3.4. The surfaces obtained by re-plumbing F along V and V 0 are isotopic. A GEOMETRIC PROOF OF THE FLYPING THEOREM 13

Figure 6. Negative (left) and positive (right) com- ponents of ∂V ∩ L. Colors: V , U, L, F \ U, ∂νL, ∂V ∩ ∂νL, ∂V ∩ F = ∂U \\∂νL, ∂U ∩ ∂νL.

Figure 7. Any arc in the boundary of a plumbing cap V joins arcs of V ∩ ∂νL with opposite signs, in this case −1 on the left and +1 on the right.

The first bulleted property in Definition 3.2 implies that if V is an acceptable plumbing cap with shadow U then (∂U ∪ ∂V ) ∩ ∂νL is a system of circles on ∂νL, each isotopic to a meridian. Consider one such circle α ∪ β, where α is an arc of ∂V ∩ ∂νL and β an arc of ∂U ∩ ∂νL. Since β ⊂ ∂F is transverse to each meridian on ∂νL, we may assign opposite local orientations to β and πL(β) ⊂ L. Extend these local orientations to global orientations on L. Convention 3.5. With the setup above, sign(α) = lk(L, α ∪ β); α is called positive if sign(α) = +1 and negative if sign(α) = −1.

Proposition 3.6. For any acceptable plumbing cap V , the arcs of ∂V ∩ ∂νL alternate in sign around ∂V . Proof. Consider any arc c of ∂V ∩ F . Orient c so that U ∩ νc lies to the left of c when viewed from V ∩ νc. Then the initial point of c has sign +1 and the terminal point has sign −1. See Figure 7.  A plumbing cap V is reducible if there is a properly embedded disk X ⊂ S3 \\(νL ∪ F ∪ V ) such that ∂X = α ∪ β for arcs α ⊂ U, β ⊂ V , and both disks V1,V2 obtained by surgering V along X are (non-fake) plumbing caps for F ; X is a reducing disk. If no such X exists, then V is irreducible. See Figure 8. 14 THOMAS KINDRED

Figure 8. A reducing disk for a reducible plumbing cap; the arc ρ in the proof of Lemma 3.7 and the neighborhood Y in the proof of Lemma 3.9.

Lemma 3.7. If V is an acceptable, irreducible plumbing cap for F with shadow U, and if ε is a properly embedded arc in U with neither endpoint on ∂νL which is parallel in S3 \\(F ∪ νL ∪ V ) to an arc ε0 ⊂ V , then ε is parallel in U to ∂V . Proof. Let W be a properly embedded disk in S3 \\(F ∪νL∪V ) with 0 ∂W = ε ∪ ε . Surger V along W to obtain two disks V1,V2; since V is irreducible, we may assume wlog that V1 is a fake plumbing cap. Denote the shadow of each Vi by Ui. If ∂V1 ⊂ F , then ε is parallel through U1 to ∂F . Assume otherwise. Then, because V1 is fake, Observation 3.3 implies that either some arc of ∂V1 ∩ ∂νL is parallel in ∂νL to ∂F , or some arc ρ of ∂V1 ∩ F is parallel in F \\U1 to ∂F . Yet, V is acceptable, and ∂V contains all arcs of ∂V1 ∩ ∂νL and all but one arc of ∂V1 ∩ F . Therefore, the one arc ρ of ∂V1 ∩ F which ∂V does not contain must cut off a disk from F \\U. That disk, however, must contain U2, implying that V2 is also a fake plumbing cap. This implies, contrary to assumption, that V was a fake plumbing cap as well.  Proposition 3.8. If X is a reducing disk for an acceptable plumbing cap V , then the disks V1,V2 obtained by surgering V along X both 18 satisfy |Vi ∩ F | < |V ∩ F |. Moreover, the surface obtained by re- plumbing along V is isotopic to the surface obtained by re-plumbing along V1 and then along V2. Proof. Let ∂X = α ∪ β, where α ⊂ U and β ⊂ V . If necessary, perturb X so that neither point of ∂α = ∂β lies on ∂F . Then |V1 ∩ F | + |V2 ∩ F | = |V ∩ F |. Moreover, both |Vi ∩ F | > 0, since α is not parallel in U to ∂F . This confirms the first statement Next, let Ui denote the shadow of Vi for i = 1, 2, and let Y be a bicollared neighborhood of X in S3 \\(F ∪νL∪V ), so that ∂Y ∩F =

18Here and throughout, | − | counts the connected components of −. A GEOMETRIC PROOF OF THE FLYPING THEOREM 15

Figure 9. A link near a crossing ball (center) with H− (left) and H+ (right).

να is a regular neighborhood in U of α and ∂Y ∩ V = νβ is a regular circ neighborhood in V of β. Note that ∂Y \ ν (α ∪ β) consists of two disks, X1 ⊂ V1 and X2 ⊂ V2. See Figure 8. Re-plumb F along V1 and then V2 by replacing U1 with V1 and then U2 with V2; next, push the disk X1 ∪ να ∪ X2 through Y to νβ. These two moves together replace U1 ∪ U2 ∪ να = U with ((V1 ∪ V2) \ \(X1 ∪ X2)) ∪ νβ = V , the same as re-plumbing F along V .  Together, Observation 3.4 and Proposition 3.8 immediately imply: Lemma 3.9. If F and F 0 are plumb-related, then there is a sequence of re-plumbing and isotopy moves taking F to F 0, in which each re- plumbing move is along an acceptable, irreducible plumbing cap.

4. Crossing balls and plumbing caps This section uses the crossing ball structures introduced in [7] to study plumbing caps for chessboards from alternating link diagrams. 4.1. Crossing ball setup. Let D be a reduced diagram (not neces- sarily alternating) of a prime nonsplit link L with crossings c1, . . . , cn, and let νS2 be a neighborhood of S2 with projection π : νS2 → S2.19 ◦ 2 Insert disjoint closed crossing balls Ct into νS , each centered at the Fn respective crossing point ct. Denote C = t=1 Ct. Perturb each Ct 2 near its equator ∂Ct ∩ S so that, for the balls Y+,Y− comprising 3 2 S \\(S ∪ C), each π|∂Y± is a diffeomorphism. Construct a (smooth) embedding of L in (S2 \ int(C)) ∪ ∂C by perturbing the arcs of D ∩ C following the over-under information at the crossings, while fixing D∩S2\int(C). Call the arcs of L∩S2 edges, and those of L ∩ ∂C ∩ ∂Y± overpasses and underpasses, respectively. Near each crossing, this looks like Figure 9, center. ◦ Let νL be a closed regular neighborhood of L in νS2 with pro- jection map πL : νL → L. For each edge e ⊂ L, call the cylinder −1 E = πL (e) ∩ ∂νL an edge of ∂νL; call the rectangles E ∩ Y± the top and bottom of the edge, respectively. For each over/underpass e± of −1 L, call E± = πL (e±)∩∂νL an over/underpass of ∂νL. Call E+ ∩Y+

19The assumption that L is prime implies that L is nontrivial. 16 THOMAS KINDRED and E+\\Y+ the top and bottom of the overpass, respectively, and call E− ∩ Y− and E− \\Y− the bottom and top of the underpass, respec- −1 tively. Assume that the meridia comprising πL (L ∩ ∂C ∩ S+ ∩ S−) −1 also comprise ∂νL∩π (∂C ∩S+ ∩S−). Then these meridia cut ∂νL into its edges, overpasses, and underpasses. 3 2 Denote the two balls of S \\(S ∪ C ∪ νL) by H±, so that each ◦ H± = Y± \ νL. Also denote ∂H± = S±. See Figure 9. −1 ◦ For each crossing ct, denote the vertical arc π (ct) ∩ Ct \ νL by S 2 ◦ vt, and denote v = t vt. For each t, ∂Ct ∩ S \ νL consists of four arcs on the equator of ∂C. Two of these arcs lie in black regions of S2 \ D, two in white. For the two arcs α, β in black regions, a core circle in α ∪ β ∪ (∂νL ∩ Ct) bounds a disk Bt ⊂ Ct such that π(Bt) is 2 disjoint from the white regions of S \ D and intersects D only at ct. This disk Bt contains the vertical arc vt and is called the standard positive crossing band at Ct. The arcs in the white regions likewise give rise to a standard negative crossing band Wt in Ct; note that ◦ Bt ∩Wt = vt. Any properly embedded disk in Ct \νL which contains ◦ vt and is isotopic in Ct \νL to Bt (resp. Wt) is called a positive (resp. negative) crossing band. See Figure 2. Denote the union of the black (resp. white) regions of S2 \ D by Bb (resp. Wc). Then   [ (7) B = Bb \ int(C ∪ νL) ∪ Bt t and   [ (8) W = Wc \ int(C ∪ νL) ∪ Wt t are the black and white chessboards from D.20 3 Denote the two balls of S \\(B ∪ W ∪ νL) by Hd±, such that each Hd± ⊃ H±; also denote ∂Hd± = Sc±. 4.2. Plumbing cap setup. Keeping the setup from §4.1, assume that D is alternating and V is an acceptable plumbing cap for B. Definition 4.1. V is in standard position if: • V is transverse in S3 to B, W, ∂C, v, and ∂V ∩ ∂νL is trans- verse in ∂νL to each meridian on ∂νL;

20Note that B ∩ W = v. Also: 2 S+ ∩ S− = S \ int(C ∪ νL) = (B ∪ W ) \ int(C), and 2 ◦ S+ ∪ S− = (S \ int(C ∪ νL)) ∪(∂C \ νL) ∪ (∂νL \ int(C)). | {z } =(B∪W )\int(C) A GEOMETRIC PROOF OF THE FLYPING THEOREM 17

Figure 10. Each arc of ∂V ∩ ∂νL disjoint from W is negative and traverses exactly one over/underpass.

• no arc of ∂V ∩ B \\∂C is parallel in B to ∂C; 21 • ∂V ∩ ∂νL ∩ C = ∅; and • W intersects V only in arcs, hence cuts V into disks. Let us check that it is possible to position V in this way. The first three conditions are straightforward. Achieve the fourth condition by removing any simple closed curves (“circles”) of V ∩ W one at a time through the following procedure. Choose a circle γ of V ∩W which bounds a disk V 0 of V \\W . Since W is incompressible, γ also bounds a disk W 0 ⊂ W . Choose a circle 0 0 00 0 0 γ ⊂ W ∩V which bounds a disk W of W \\V . (If int(W )∩V = ∅, then γ0 = γ.) The circle γ0 also bounds a disk V 00 ⊂ V . The sphere V 00 ∪ W 00 bounds a ball Y in the link exterior, since L is nonsplit. Push V 00 through Y past W 00, removing γ0 from V ∩ W . Repeat this process until W intersects V only in arcs, and therefore cuts V into disks. Observation 4.2. If V is in standard position, then all components of V ∩ Hd± are disks, and all components of Hd± \\V are balls.

Observation 4.3. If V is in standard position, then each arc of ∂V ∩ ∂νL which is disjoint from W is a negative arc that traverses exactly one overpass or underpass (along the top or bottom, respectively). This follows from Convention 3.5 and the fact that D is alternat- ing; see Figure 10. Observation 4.3 implies, in particular: Observation 4.4. Any positive arc of ∂V ∩ ∂νL lies entirely on a single edge of ∂νL and contains an endpoint of an arc of V ∩ W .

Proposition 4.5. V ∩W 6= ∅, and each arc α of ∂V \\W intersects at most two disks of B \\W .

21That is, ∂V is disjoint from the top of each overpass and bottom of each underpass of ∂νL. Hence, for any overpass that ∂V intersects, ∂V traverses the entire top of the overpass; likewise for underpasses of ∂νL. 18 THOMAS KINDRED

Figure 11. The possible configurations for V near an innermost circle of V ∩ Sc+.

Proof. Observation 4.4 and Proposition 3.6 imply that V ∩ W 6= ∅. Consider an arc α of ∂V \\W . Proposition 3.6 and Observation 4.4 imply that α intersects at most one negative arc of ∂V ∩ ∂νL. Also, each arc of α \\∂νL lies in a single disk of B \\W . The result now follows from Observations 4.3 and 4.4. 

Next, consider an outermost disk V0 ⊂ V \\W . Its boundary consists of an arc α ⊂ ∂V \\W and an arc β ⊂ V ∩ W . Let να be a regular neighborhood of α in ∂V with να ∩ ∂B ⊂ α.

Lemma 4.6. If such V0 ⊂ Hd+, then V0 appears as in Figure 11. In particular, the endpoints of να lie in opposite disks of Sc+ \\∂V0.

Proof. Let ∂V0 = α ∪ β as above. Since V is in standard position, the endpoints of α cannot lie on the same vertical arc or edge of ∂νL. This and the fact that D is prime and reduced imply that α must traverse an overpass. Hence, by Proposition 4.5, α intersects exactly two disks of B \\W , and its configuration depends only on its endpoints, each of which lies either on v or on a positive arc of ∂V ∩ ∂νL. Figure 11 shows the three possibilities. 

With V0 as in Lemma 4.6: Observation 4.7. There is exactly one arc δ of ∂V \\W whose endpoints lie in opposite disks of Sc+ \\∂V0. This arc δ traverses the underpass at the same crossing where ∂V0 traverses the overpass. 4.3. Plumb-equivalence and flype-equivalence. Maintain the setup from §§4.1-4.2. In particular, D is a reduced alternating dia- gram of a prime nonsplit link L; B and W are the chessboards from D; and a plumbing cap V for B is in standard form. Say that V is apparent in D if |V ∩ W | = 1. Lemma 4.6 implies that any apparent plumbing cap for B appears as in Figure 12, left. Proposition 4.8. If V is apparent, then V is irreducible. Proof. If V were reducible, then it would be possible to surger V to give two plumbing caps V1,V2 with both |∂Vi∩∂νL| < |∂V ∩∂νL| = 4. A GEOMETRIC PROOF OF THE FLYPING THEOREM 19

Figure 12. Every plumbing cap for W is either iso- topically apparent (left) or reducible (right).

Since both |∂Vi ∩ ∂νL| must be even, they must both equal two. But then V1, V2 would each decompose F as a boundary connect sum. This is impossible because F is essential and L is prime.  Theorem 4.9. If V is an acceptable, irreducible plumbing cap for B in standard position, then V is apparent in D.

Proof. Proposition 4.5 implies that V ∩ W 6= ∅. Assume for con- tradiction that |V ∩ W | > 1. Then there is a disk V1 ⊂ V whose boundary consists of two arcs, α ⊂ ∂V and β ⊂ V ∩ W , and whose interior intersects W in a nonempty collection of arcs β1, . . . , βk, such that each βi is parallel through a disk Xi ⊂ V1 to an arc αi ⊂ α. Assume wlog that the disks Xi all lie in Hd−. Denote γ = α ∪ β = ∂V1, and for each i = 1, . . . , k, denote the circle αi ∪ βi = ∂Xi by γi. Then γ1, . . . , γk are disjoint simple closed curves in Sc+. Relabel if necessary so that γ1 bounds a disk Y ⊂ Sc+ 22 0 which is disjoint from β∪γ2 ∪· · ·∪γk. Let δ, δ be the arcs of ∂V \\W 0 that each share an endpoint with α1. The other endpoints of δ, δ 23 must both lie in the disk Sc+ \\Y . Lemma 4.6 implies wlog that the endpoints of δ lie in opposite disks of Sc+ \ ∂γ1. Hence, Observation 4.7 implies that δ traverses the underpass at the same crossing where γ1 traverses the overpass. Moreover, ∂V is disjoint from the vertical arc at this crossing, or else an arc of ∂V \\W would have both endpoints on that vertical arc. Therefore, there is a properly embedded arc ε in the shadow U of V which is disjoint from W and has one endpoint on δ ∩ int(B) and the other on δ0 ∩ int(B), as in Figure 12, right. Further, since δ and 0 δ both lie on the boundary of the disk V1 \\(X1 ∪· · ·∪Xk) ⊂ V \W ,

22 This is possible because at least two circles among γ1, . . . , γk bound disks of Sk Sc+ \\ i=1 γi, and most one of these disks intersects β. 23 This is because these endpoints lie on β ∪ γ2 ∪ · · · ∪ γk ⊂ Sc+ \\Y . 20 THOMAS KINDRED

T T 2 2 2 T

T T 1 1

Figure 13. A flype move corresponds to an iso- topy of one spanning surface (not shown) and a re- plumbing of the other (shown). that disk contains an arc ε0 with the same endpoints as ε. The arcs ε 0 and ε lie on the boundary of the same ball of Hd−.Thus, ε is parallel in S3 \\(F ∪ νL ∪ V ) to the arc ε0 ⊂ V , and so Lemma 3.7 implies that ε is parallel in U to one of the two arcs of ∂V \\∂ε. The arc of ∂V \\∂ε which contains α1 is not a subset of ∂U, and so ε must be parallel in U to the other arc of ∂V \\∂ε. But then that arc is disjoint from W and so |V ∩ W | = 1, contrary to assumption.  Given an apparent plumbing cap V for B (or W ), there is a cor- responding flype move, as shown in Figure 13. Namely, the flype move proceeds along a circle γ ⊂ S2 comprised of the arc V ∩ W together with an arc in U ∪ νL. (The resulting link diagram might be equivalent to D.) Conversely, if D → D0 is a flype move along a curve γ ⊂ S2, then there is an apparent plumbing cap V for W or B with V ∩ S2 ⊃ γ \\νL. Proposition 4.10. Suppose that V is an apparent plumbing cap for B, D → D0 is the flype move corresponding to V , B0 and W 0 are the chessboards from D0, and B00 is the surface obtained by re-plumbing B along V . Then B0 and B00 are isotopic, as are W 0 and W .24 Proof. Figure 13 demonstrates the isotopies.  Theorem 4.11. Let D and D0 be reduced alternating diagrams of a prime, nonsplit link L with respective chessboards B,W and B0,W 0. Then D and D0 are flype-related if and only if B and B0 are plumb- related as are W and W 0. Proof. If D and D0 are flype-related, then B and B0 are plumb- related as are W and W 0, by Proposition 4.10. Assume conversely that B and B0 are plumb-related, as are W and W 0. Then, by Lemma 3.9, there are two sequences of re-plumbing 0 0 moves B = B0 → · · · → Bm = B and W = W0 → · · · → Wn = W , such that each re-plumbing move Bi → Bi+1 (resp. Wi → Wi+1) 0 follows an acceptable, irreducible plumbing cap Vi (resp. Vi ). Let D0 = D. Put the plumbing cap V0 in standard position with respect to D. Theorem 4.9 implies that V0 is now apparent. Let

24An analogous statement holds for plumbing caps for W . A GEOMETRIC PROOF OF THE FLYPING THEOREM 21

D0 → D1 be flype move corresponding to V0. By Proposition 4.10, the negative-definite chessboard from D1 is isotopic to W , and the positive-definite chessboard from D1 is isotopic to B1. Repeat in this manner. Ultimately, this gives a sequence of fly- pes from D = D0 to a diagram Dm of L whose positive-definite chessboard is isotopic to B0 and whose negative-definite chessboard is isotopic to W . To complete the proof, construct a sequence of 0 0 0 flypes Dm = D0 → · · · → Dn = D in the same manner, but now with the flypes coming from the (apparent) acceptable, irreducible 0 plumbing caps Vi for Wi.  Corollary 4.12. If F is an essential positive-definite surface span- ning an alternating link, and F 0 is plumb-related to F , then F 0 is also essential and positive-definite.

5. Spanning surfaces in the crossing ball setting

Adopt the setup from 4.1 of D, B, W , C, νL, H±, S±, etc. Now, also let F be an incompressible spanning surface for a prime nonsplit link L. Keep this setup throughout §5. The opening of each subsection of §5 will declare any additional hypotheses for that subsection. Starting in §5.4, these hypotheses will become increasingly specific regarding F and L. Namely: • Starting in §5.4, F is essential. • Starting in §5.5, L and D are alternating, hence B is positive- definite, and W is negative-definite. • Starting in §5.6, F is definite. 5.1. Fair position for F . Definition 5.1. F is in fair position if: • F is transverse in S3 to B, W , ∂C, and v; • ∂F is transverse in ∂νL to each meridian on ∂νL; • whenever Ct ∩ ∂F 6= ∅, Ct contains a crossing band in F ; • each crossing band in F is disjoint from S+; and • S+ ∪ S− cuts F into disks. (Later, we will define a slightly more restrictive good position for F .) Let us check that it is possible to position F in this way. The first four conditions are straightforward. (The fourth condition implies that every crossing band in F appears as in Figure 14.) One way to achieve the fifth condition is by minimizing |F \ (S+ ∪ S−)|, subject to the first four conditions.25 25 Suppose that |F \ (S+ ∪ S−)| is minimized, and assume for contradiction that some component X of F \\(S+ ∪ S−) is not a disk. Choose a component of ∂X and push it into int(X); the resulting circle γ is 0-framed in F and does not 3 bound a disk in X. Since γ is 0-framed, it bounds a disk in S \ (S+ ∪ S− ∪ νL). Among all such disks, choose one, Z, which intersects F transversally (hence in 22 THOMAS KINDRED

Figure 14. Positive (left) and negative (right) cross- ing bands in a surface F in fair (or good) position.

Proposition 5.2. If F is in fair position, then all components of C \\F and H± \\F are balls, and all components of F ∩ S+ ∩ S−, F ∩ ∂C ∩ S±, and F ∩ ∂νL ∩ S± are arcs. Proof. By assumption, F intersects each crossing ball in disks, hence it cuts these crossing balls into balls. Likewise for H±. All components of ∂C ∩ S+, ∂C ∩ S−, and S+ ∩ S− are disks. If F intersected one of these disks in a circle, γ, then, since all components of F \\(S+ ∪ S− ∪ νL) are disks, γ would bound disks of F in both components of F \\(S+ ∪S− ∪νL) whose boundaries contain γ. This contradicts the assumption that F is connected. The assumption that D is nontrivial and nonsplit implies that each component of F ∩ ∂νL ∩ S± is an arc.  5.2. Bigon Moves.

Definition 5.3. Let F be in fair position and α ⊂ S± an arc that: • intersects F precisely on its endpoints, which lie on the same circle γ of F ∩ S±, but not on the same arc of F ∩ S+ ∩ S−; • is disjoint from over/underpasses; and • intersects S+ ∩ S− in exactly one component. Then α is parallel through a properly embedded (“bigon”) disk Z ⊂ H± \\F to an arc β ⊂ F ∩ H±, and one can push F ∩ νβ through νZ past α, as in Figures 15 and 16. This is called a bigon move.

Proposition 5.4. Given an arc α ⊂ S± as in Definition 5.3, the endpoints of α do not lie on the same circle of F ∩ S∓. simple closed curves only) and in the fewest possible number of circles. Choose an innermost disk Z0 ⊂ Z \\F , and denote ∂Z0 = γ0. The incompressibility of F implies that γ0 bounds a disk X0 in F . Then X0 ∪ Z0 is a disk which bounds a 3 ball Y ⊂ S \\νL; the interior of Y is disjoint from F , because the interior of Z0 is disjoint from F . Yet, X0 must intersect S+ ∪S−, due to the minimality of |Z ∩F |. Push F near X0 through Y past Z0, thus decreasing |F \ (S+ ∪ S−)|. Perform this isotopy in such a way that the resulting positioning satisfies the first three conditions (this is possible because X0 must be disjoint from any crossing bands, and it must contain any components of F ∩ C that it intersects). Contradiction. A GEOMETRIC PROOF OF THE FLYPING THEOREM 23

Figure 15. The three types of bigon moves; the mid- dle one also appears in Figure 16.

β

α

Figure 16. One of three types of bigon moves.

Proof. If either endpoint of α is on ∂νL, instead of in S+ ∩ S−, there is nothing to prove. Assume that α ⊂ S+ ∩ S−, and assume for contradiction that γ±, respectively, are circles of F ∩ S±, both of which contain both endpoints of α. Let X±, respectively, denote the disks of F ∩ H± with ∂X± = γ±. Choose arcs β± ⊂ X± with the same endpoints as α, and let δ± = α ∪ β±. The circles δ± bound properly embedded disks Z± ⊂ H± \\F . Gluing Z+ and Z− along α produces a disk Z with Z ∩ F = ∂Z = β+ ∪ β−. Since F is incompressible, β+ ∪ β− must bound a disk X ⊂ F . The boundary of this disk X is the curve β+ ∪β−, which intersects S+∩S− only in two points. Therefore, X intersects S+∩S− in a single arc, and this arc has the same endpoints as α. Contradiction.  Lemma 5.5. Performing a bigon move on a surface in fair position produces a surface in fair position. Proof. Consider a bigon move along an arc α. Assume wlog that the endpoints of α lie on the same circle γ+ of F ∩ S+. Then, by 0 Proposition 5.4, the endpoints of α lie on distinct circles γ−, γ− of F ∩ S−. Let X+ be the disk of F ∩ H+ bounded by γ+, and let X− 0 0 and X− be the disks of F ∩ H− bounded by γ− and γ−, respectively. 0 Since X− 6= X−, the bigon move along α must involve pushing an arc β+ ⊂ X+ past α. This has the effect of splitting X+ into two 0 distinct disks of F ∩H+, while merging X− and X− into a single disk of F ∩ H−. The rest of F is unaffected. Thus, this move preserves the fact that S+ ∪ S− cuts F into disks. Hence, it produces a surface in fair position.  5.3. Good position for F . Definition 5.6. F is in good position if: 24 THOMAS KINDRED

Figure 17. Removing arcs of ∂F ∩ ∂νL ∩ S± with both endpoints on the same arc of S+ ∩ S− ∩ ∂νL.

Figure 18. Removing arcs of F ∩∂C ∩S± with both endpoints on the same arc of ∂C ∩ S+ ∩ S−.

• F is in fair position, • no arc of ∂F ∩ ∂νL ∩ S± has both endpoints in the same arc of S+ ∩ S− ∩ ∂νL, and • no arc of F ∩ ∂C ∩ S± has both endpoints on the same arc of ∂C ∩ S+ ∩ S−. Given a surface F in fair position, it is possible to use the moves in Figures 17 and 18 to remove any arcs described in Definition 5.6, all while preserving the conditions of fair position. That is, any surface in fair position admits local adjustments that put it in good position. Proposition 5.7. If F is in good position, then each component of F ∩ C is either a crossing band or a saddle disk as in Figures 14, 19.

Proof. Consider a crossing ball Ct where F does not have a crossing band. Each component γ of F ∩ Ct is a circle because F does not have a crossing band at Ct, and so, by assumption, ∂F ∩ ∂Ct = ∅. Also, by assumption, each such circle γ bounds a disk of F ∩ Ct. Moreover, Proposition 5.2 implies that S2 cuts each such γ into arcs; by assumption, the endpoints of each arc of γ \\S2 are on A GEOMETRIC PROOF OF THE FLYPING THEOREM 25

Figure 19. When F is in good position, F ∩ C is comprised of crossing bands and saddle disks.

Figure 20. Left: If an arc β of ∂F ∩ S− is disjoint from underpasses, then its endpoints lie on distinct circles of F ∩ S+. Right: In particular, no arc of F ∩ S+ ∩ S− is parallel in S+ ∩ S− to ∂νL. distinct arcs of ∂Ct ∩ S+ ∩ S−. Since each disk of ∂Ct ∩ S± contains only two arcs of of ∂Ct ∩ S+ ∩ S−, the result follows. 

5.4. Essential spanning surfaces. Assume in §5.4 that F is an essential surface in good position. Define the complexity of F to be (9) ||F || = |v \\F |. Since F is in good position, Proposition 5.7 implies that any disk R of S+ ∩ S− and any crossing ball Ct incident to R satisfy: ( 0 F has a crossing band at Ct (10) |vt \\F | = |∂R ∩ ∂Ct ∩ F | otherwise.

Proposition 5.8. For each arc α of ∂F ∩ ∂νL ∩ S± that traverses no overpasses, the endpoints of α lie on distinct circles of F ∩ S∓. Proof. Assume for contradiction that the endpoints of such an arc α lie on the same circle γ of F ∩ S∓, and let X be the disk of F ∩ H∓ bounded by γ. Then X contains a properly embedded arc β with the same endpoints as α. The fact that α traverses no overpasses 3 0 implies that β is parallel in S \\νL to an arc β on ∂νL ∩ S∓. Since 26 THOMAS KINDRED

Figure 21. No arc of ∂F ∩ S+ ∩ S− has both end- points on the same crossing ball.

F is in good position, the shared endpoints of these arcs must lie on 0 distinct disks of S+ ∩ S−. Therefore, the circle α ∪ β is a meridian on ∂νL. Hence, β0 is not parallel in ∂νL to ∂F . This is impossible, by Proposition 2.14. 

Proposition 5.9. No arc of F ∩ S+ ∩ S− is parallel in S+ ∩ S− to ∂νL. Proof. If such an arc exists, then choose one, α which is outermost in S+ ∩ S−. Up to mirror symmetry, α must appear as in Figure 20, right. This is impossible, by Proposition 5.8. 

Proposition 5.10. If |v\\F | is minimal, then no arc of ∂F ∩S+∩S− has both endpoints on the same crossing ball.

Proof. If both endpoints of an arc of ∂F ∩ S+ ∩ S− lie on the same crossing ball, then there is an outermost such arc α in S+ ∩S−. Both endpoints of α lie on saddle disks, by 5.7, so α abuts two saddle disks. The sequence of isotopy moves shown in Figure 21 decreases |v \\F |, so it was not minimal. 

Proposition 5.11. If |v \\F | is minimized and a circle γ of F ∩ S± traverses the over/underpass at a crossing ball Ct, then γ ∩∂Ct = ∅.

Proof. Assume for contradiction that γ traverses the overpass at Ct and γ ∩ ∂Ct 6= ∅. Then there is a disk R of S+ ∩ S− containing a point x of γ ∩ ∂Ct ∩ S+ ∩ S−. Also, as shown in Figure 22, there is an arc α ⊂ S+ disjoint from overpasses with α ∩ S+ ∩ S− ⊂ R and ∂α ⊂ γ; Proposition 5.10 implies that such may position α with α ∩ F = ∂α.26 Following Figure 22, perform a bigon move along α, and then perform an additional isotopy move, which removes a saddle disk and thus decreases |v \\F |, contrary to assumption.  In particular, Proposition 5.11 implies the following two facts: Proposition 5.12. If |v \\F | is minimized, then no arc α of F ∩ S+ ∩ S− has one endpoint on C and the other on an incident edge of ∂νL.

26The proof of Proposition 5.14 employs a similar argument and provides greater detail on this point. A GEOMETRIC PROOF OF THE FLYPING THEOREM 27

Figure 22. If |v \\F | is minimized and a circle γ of F ∩ S+ traverses the overpass at Ct, then γ ∩ Ct = ∅.

Proof. If such an arc exists, choose one, α, which is outermost in S+ ∩ S−. Let x denote the endpoint of α on a crossing ball Ct, and assume wlog that the edge of ∂νL containing the other endpoint y of α abuts the overpass at Ct. Consider the arc β of ∂F on this edge between y and the overpass at Ct. If β∩S+∩S− = ∅, the result follows immediately from Proposition 5.11. Otherwise, by Proposition 5.9 and the fact that α is outermost, β intersects S+ ∩S− in exactly one point. In this case, there is now a bigon move which causes the circle of F ∩ S+ traversing the overpass 27 at Ct to intersect ∂Ct. This contradicts Proposition 5.11.  Proposition 5.13. If |v \\F | is minimized, then for each crossing ball Ct where F does not have a crossing band, the arcs of F ∩∂Ct∩S± all lie on distinct circles of F ∩ S± from one another and from the arc of ∂F ∩ S± ∩ ∂νL which traverses the over/underpass at Ct.

Proof. Otherwise, a bigon move near Ct gives an immediate contra- diction to Proposition 5.10 or Proposition 5.12.  Proposition 5.14. If |v \\F | is minimized and a circle γ of F ∩ S± traverses the over/underpass of ∂νL at a crossing Ct, then γ is disjoint from both edges of ∂νL incident to the under/overpass at Ct.

Proof. Assume for contradiction that a circle γ of F ∩ S+ traverses the overpass of ∂νL at Ct and intersects an edge of ∂νL incident to the underpass at Ct. Then this edge of ∂νL contains an endpoint of an arc of γ ∩ S+ ∩ S−. Denote the disk of S+ ∩ S− containing this endpoint by R. As shown in Figure 23, there is an arc α ⊂ S+ disjoint from overpasses with α ∩ S+ ∩ S− ⊂ R and ∂α ⊂ γ. Choose

27 To see the bigon move, let ρ be the arc of F ∩ S+ ∩ S− with an endpoint on 0 int(β); let α be the arc of F ∩ ∂Ct ∩ S− sharing an endpoint with α; and let σ be 0 the arc of F ∩ S+ ∩ S−, other than α that shares an endpoint with α . The bigon move is along an arc in S+ ∩ S− with one endpoint on ρ and the other on σ. 28 THOMAS KINDRED

Figure 23. No circle of F ∩ S+ intersects both the overpass at Ct and an edge of ∂νL incident to the underpass at Ct.

α such that α t F and all points of int(α) ∩ F lie on distinct arcs of F ∩ S+. We claim that α is properly embedded in S+ \\F , i.e. α∩F = ∂α. Suppose otherwise. Then there is an arc α0 of α \\F sharing neither endpoint with α, such that both endpoints of α0 lie on the same circle 0 of F ∩S+. Perform a bigon move along α . Consider the resulting arc β that lies in the disk of R \\(γ ∪ α) incident to Ct. Each endpoint of β lies either on the arc ∂Ct ∩ R or on the arc of ∂νL ∩ R incident to the underpass at Ct. This is impossible, by Propositions 5.10, 5.9, and 5.12. Therefore, α is properly embedded in S+ \\F . Following Figure 23, perform a bigon move along α, and then adjust F to have a crossing band at Ct. This decreases |v \\F |, contrary to assumption.  5.5. Alternating links. Assume again in §5.5 that F is an essential surface in good position. Assume also that its complexity ||F || = |v \\F | is minimized, and that D is a reduced alternating diagram (of a prime nonsplit link L) with chessboards B and W .28 Much of the work in §5.5 will address an innermost circle γ of F ∩ S+, in the following setup. Let T+ be an innermost disk of S+ \\F , and denote ∂T+ = γ. Let X be the disk that γ bounds in 0 29 F , and let T+ = S+ \\T+. Denote −1 T− = S− ∩ π (π(T+)) . Before addressing γ, we note two more general consequences of the assumption that D is alternating: Proposition 5.15. F ∩ B and F ∩ W contain no ∂-parallel arcs.

28Recall Convention 2.3. 29Recall that π denotes the projection map νS2 → S2. A GEOMETRIC PROOF OF THE FLYPING THEOREM 29

Figure 24. The situation at the end of the proof of Proposition 5.15, with α,α 0,X, and β.

Proof. Assume for contradiction that an arc α of F ∩B is ∂-parallel in S3 \\νL. Then, since D is reduced, B is essential, and so α cuts off a disk X from B with ∂X = α∪α0 for some arc α0 ⊂ ∂B; good position implies that X ∩ W 6= ∅. Let Z be an outermost disk of X \\W , so that ∂Z = β ∪ β0, where β ⊂ X ∩ W ⊂ v and β0 ⊂ ∂X = α ∪ α0. If β0 ⊂ α, then α ⊂ F ∩ B is parallel in B to v. Then, pushing F near α through Z past β reduces |v \\F |, contrary to assumption. If β0 ⊂ α0, then α0 ⊂ ∂B0 is parallel in B to v. This is impossible, because D is reduced. Otherwise, β0 contains one of the shared endpoints x of α, α0; as shown in Figure 24, it is possible to push α through Z past β ⊂ v, while sliding x along α0 ⊂ ∂B, in a way the decreases |v \\F |, contrary to assumption.  The assumptions that that D is alternating and F is in good po- sition imply: Observation 5.16. Near each edge of ∂νL, F appears as in Figure 25 (up to mirror symmetry). Note in particular that an edge of ∂νL contains an odd number of points of F ∩ S+ ∩ S− if and only if ∂F traverses the top of the overpass incident to that edge. We now turn our attention to the innermost circle γ of F ∩ S+. Proposition 5.17. Any arc of γ ∩ ∂C appears as in Figure 26.

Proof. Suppose γ intersects a crossing ball Ct. Then Proposition 5.13 implies that γ must intersect Ct in one arc only and must not traverse the overpass at Ct. Proposition 5.12 further implies that γ is disjoint from the edge E of ∂νL which abuts both the underpass at Ct and 2 the disk of ∂Ct \(S ∪νL) incident to γ. Hence, E ∩F ∩S+ ∩S− = ∅, and so Observation 5.16 implies that F must have a crossing band at the opposite end of this edge.  30 THOMAS KINDRED

Figure 25. How F appears near each edge of ∂νL in §5.5 (up to mirror symmetry) and §§5.6-5.7 (no reflection allowed).

Proposition 5.18. Each arc of γ∩S+ ∩S− has at most one endpoint on C. Proof. This follows immediately from Proposition 5.17 and the fact that D is alternating. 

Say that γ encloses a crossing ball Ct if π(T+) ⊃ π(Ct). Lemma 5.19. F has a crossing band in every crossing ball which γ encloses, and γ encloses at least one crossing ball.30

Proof. If γ encloses a crossing ball Ct, then ∂F does not traverse the overpass at Ct, since γ is an innermost circle of F ∩ S+. Initial position implies that F has a crossing band at Ct. Assume, contrary to the second claim, that γ encloses no crossing balls. Then γ cannot intersect a crossing ball, due to Proposition 5.17. Further, γ cannot traverse an overpass, because of Proposition 5.14 and Observation 5.16. Therefore, T+ is disjoint from all crossing balls and overpasses. Hence, π(γ) bounds a disk in S2 \ C. This implies, contrary to assumption, that some arc of γ ∩ S+ ∩ S− is parallel in S+ ∩ S− to ∂νL. 

5.6. Definite surfaces. Assume again in §5.6 that F is an essential surface in good position, ||F || = |v \\F | is minimized, and D is a reduced alternating diagram of L with chessboards B and W . Also maintain the setup from §5.5 of the innermost circle γ ⊂ F ∩ S+ and

30This gives an alternate proof of the crossing band lemma from [1]: Crossing Band Lemma from [1]. Given a reduced alternating diagram D of a prime nonsplit link L and an essential spanning surface F for L, it is possible to isotope F so that it has a crossing band at some crossing in D. A GEOMETRIC PROOF OF THE FLYPING THEOREM 31

Figure 26. The vicinity of an arc of γ ∩∂C. Inward- pointing arrows along the innermost circle γ of F ∩S+ identify black and white regions of S+ ∩ S−.

0 the associated disks T+,T+ ⊂ S+ and T− ⊂ S−. Now assume also that F is positive-definite. Much of the work in §5.6 will address the possibilities for arcs δ of F ∩ T−. First, we note the following more general consequences of the assumption that F is positive-definite: Lemma 5.20. Every arc α of F ∩ W satisfies

i(∂F, ∂W )ν∂α = −2. In particular, every crossing band in F is positive. Proof. Proposition 5.15 implies that α is not ∂-parallel, and so Lemma 2.16 gives i(∂F, ∂W )ν∂α = −2.  Observation 5.16 and Lemma 5.20 immediately imply: Observation 5.21. ∂F intersects each edge of ∂νL as in Figure 25.

Now consider F ∩ T−. Each component of F ∩ int(T−) is disjoint from S+ (since γ is an innermost circle of F ∩ S+) and thus consists alternately of arcs on ∂νL ∩ S− and arcs on ∂C ∩ S− which abut crossing bands in F . This and the fact (due to Lemma 5.20) that every crossing band in F is positive imply that every circle of F ∩ int(T−) encloses some disk of B \\C in T− and is parallel in (∂νL ∪ ∂C) ∩ S− to the boundary of this disk. Likewise, for each arc δ of F ∩int(T−), some disk B0 of B ∩T− \\C abuts all components of ∂νL ∩ S− and ∂C ∩ S− that intersect δ.

Convention 5.22. Orient each arc δ of F ∩ int(T−) so that this nearby disk B0 lies to the right of δ, when viewed from H+. Denote the initial point of δ by x and the terminal point by y.

Proposition 5.23. Let δ be an arc of F ∩int(T−), following Conven- tion 5.22. Then x either lies on an underpass of ∂νL or is a point in ∂νL∩B\C with i(∂F, ∂B)νx = −1. Also, y ∈ ∂νL∩S+∩S−; if y ∈ B, then i(∂F, ∂B)νy = +1; otherwise, y ∈ W with i(∂F, ∂W )νy = −1. 32 THOMAS KINDRED

Figure 27. Endpoints of an arc δ of F ∩ int(T−), oriented as in Convention 5.22.

Proof. Each endpoint of δ lies either on γ ∩ ∂νL ∩ S+ ∩ S− or on an underpass of ∂νL in π−1(π(γ)). Lemma 5.20 implies that any endpoint z of δ on ∂νL ∩ W satisfies i(∂F, ∂W )νz = −1. Thus, every endpoint of δ has one of the types described in the proposition. Moreover, because D is alternating and B0 lies to the left of δ, each type of endpoint can occur only at x or only at y as stated.  Although there are three possibilities for x and two for y, there are actually just three possibilities for the pair (x, y):

Proposition 5.24. Let δ be an arc of F ∩int(T−), as in Convention 5.22. Then x ∈ B if and only if y ∈ B. Proof. We argue by contradiction. Suppose first that x ∈ B and y∈ / B. Let γx ⊂ γ be the arc of F ∩ S+ ∩ S− that has x as an endpoint, and let γy ⊂ γ be the arc of F ∩ S+ ∩ ∂νL that has y as 0 0 an endpoint. Let x ∈ int(γx) and y ∈ int(γy). See Figure 28, left. Perform a bigon move along the arc α.31 Then consider the arc of ∂F ∩ ∂νL ∩ S+ that contains y. Both endpoints of this arc lie on the circle of F ∩ S− that contains δ, contradicting Proposition 5.8. Suppose now that x∈ / B and y ∈ B. Let γy ⊂ γ be the arc of 0 F ∩ S+ ∩ S− that has y as an endpoint, and let y ∈ int(γy). The initial point x lies on the underpass of ∂νL at a crossing Ct, and 00 00 00 00 π(x) = π(x ) for some x ∈ γ. Either x ∈ ∂νL or x ∈ ∂Ct. 00 Suppose first that x ∈ ∂νL, as in Figure 28, center. Let γx be the arc of γ on the edge of ∂νL that abuts B0 and the overpass 0 containing x, and let x ∈ int(γx). As in the first case, it is now possible to perform a bigon move along an arc α with endpoints 0 0 x and y . Do so. The circle of F ∩ S− containing δ traverses the underpass at Ct and now has an endpoint on the edge of ∂νL incident to the overpass at Ct. This contradicts Proposition 5.14. In the remaining case, a similar argument leads to a contradiction, using Proposition 5.11. See Figure 28, right. 

31 0 0 Note that x ∈ ∂B0, and y is on the same edge of ∂νL that contains y. This edge abuts B0, so there is a properly embedded arc α ⊂ S+ \\F with endpoints 0 0 x , y ∈ γ which is disjoint from overpasses and intersects S+ ∩ S− only in B0. Since y0 ∈ ∂νL, it is possible to perform a bigon move along α. A GEOMETRIC PROOF OF THE FLYPING THEOREM 33

y yx y x 0 0 0 0 0 y x y x y x x0

Figure 28. These bigon moves show that the initial point x of δ lies in B iff the terminal point y does.

Proposition 5.25. Let δ be an arc of F ∩ int(T−), following Con- vention 5.22. If x, y ∈ B, then δ is parallel in T− to ∂T−.

Proof. In this case, x, y ∈ γ ∩ B0. Thus, Proposition 5.4 implies that x and y lie on the same arc γ0 of γ0 ∩ S+ ∩ S−. Hence, δ ∪ γ0 is a circle in T− \\F , so δ is parallel in T− to γ0 ⊂ γ ∩ S− ⊂ ∂T−.  5.7. Proof of Theorem 5.26 and the Flyping Theorem. Let L be a prime nonsplit link, and let D be a reduced alternating diagram of L with chessboards B and W , which are respectively positive- and negative-definite. Theorem 5.26. Any essential positive-definite surface F spanning L is plumb-related to B. Proof. While fixing B and W , transform F by isotopy so that it is in good position and ||F || = |v \\F | is minimal. If F ∩ H+ = ∅, then F traverses the top of no overpass, hence has a crossing band at every crossing. Lemma 5.20 implies that each crossing band is positive, and so F is isotopic to B. Assume instead that F ∩ H+ 6= ∅, and let T+ be an innermost disk of S+ \ F , set up as at the beginning of §5.5, with T− = S− ∩ −1 π (π(T+)) and γ = ∂T+ = ∂X for X ⊂ F ∩ H+. Propositions 5.23, 5.24, and 5.25 imply that each arc δ of F ∩ int(T−) has one one the three types (I, II, or III from left to right) shown in Figure 29. Note:

(11) At least one arc δ of F ∩ int(T−) is of type II or III.

Choose a point z in a region of W ∩ T+ abutting γ. While fixing ◦ 2 −1 F ∩ (S+ ∪ S− ∪ C), adjust F ∩ H± so that F ⊂ νS \ π (z) and 34 THOMAS KINDRED

Figure 29. Arcs of δ ∩ int(T−), and the arcs used in the proof of Theorem 5.26 to transform γ into γ0.

32 π restricts to an injection on each disk of F ∩ H±. Then, while fixing F \\X, push X through the point at infinity in H+ and into −1 H+ ∩ π (T+) such that π : X → π(T+) bijectively. For each arc δ of F ∩int(T−) of type II or III, construct a properly −1 embedded arc in T+ \ (B ∪ π (C)), as shown in Figure 29. Denote the union of these arcs by ρ, and denote the disk of T+ \\ρ containing 0 −1 0 0 0 0 z by T+. Also denote S− ∩ π (π(T+)) = T− and ∂T+ = γ . Take an annular neighborhood A of π(γ0) in S2, such that A inter- 0 sects only the crossing balls that π(γ ) intersects, ∂A ∩ C = ∅, and 0 each arc of F ∩S+ ∩S− that intersects A lies on γ or has an endpoint on a saddle disk at a crossing where γ0 abuts a saddle disk. Denote 0 2 ∂A = γ0 ∪ γ1, such that γ0 ⊂ π(T+). Denote S \\A = Z0 t Z1, −1 where each ∂Zi = γi. Denote π (Z0) = Y0. 0 Denote the arcs of γ ∩ W by ω1, . . . , ωm. Each ωi has a dual arc 33 αi ⊂ A ∩ W \ C. Denote the rectangles of A \\(α1 ∪ · · · αm) by A1,...,Am such that each ∂Ai ⊃ αi ∪ αi+1, taking indices modulo −1 m. Denote each prism π (Ai) by Pi. Each prism Pi intersects F in one of the three ways indicated in the 34 left column of Figure 30. For each i, let Fi denote the component of F ∩ Pi which abuts γ. Each Fi is a disk. Observe that Fi and Fj intersect in an arc when i ≡ j ± 1 (mod m) and are disjoint when i 6≡ j, j ± 1 (mod m). Therefore, F1 ∪ · · · ∪ Fm is either an annulus or a mobius band; denote it by FA. Moreover, the disk X ∩ Y0 attaches to FA along its boundary. Therefore, FA is an annulus, and the following subsurface of F is a disk:

(12) U = (X ∩ Y0) ∪ FA.

32 This may involve pushing some of the disks of F ∩ H+ through the “point at infinity” in H+. 33 The arc αi has one endpoint on γ0 and one on γ1, with |ωi ∩ αi| = 1. 34 The green arcs top-left describe a disk in Pi \ νL, which is parallel through −1 a ball Y ⊂ Pi to π (γ1); F intersects Pi as shown and in an arbitrary number of additional disks in Y , each containing a saddle disk. A GEOMETRIC PROOF OF THE FLYPING THEOREM 35

Figure 30. The re-plumbing move in the proof of Theorem 5.26.

◦ There is a properly embedded disk V ⊂ νS2 \\(F ∪ νL) which intersects Y0 in a disk (in H−) and intersects each prism Pi as indi- cated in the right column of Figure 30.35 Note that ∂V ∩F = ∂U ∩F and (∂V ∩ ∂νL) ∪ (∂U ∩ ∂νL) is a system of meridia (top two rows of Figure 30) and inessential circles (bottom row of the figure) on ∂νL. Thus V is a, possibly fake, plumbing cap for F , and U is its shadow. Whether or not V is fake, Proposition 3.1 implies that F is plumb- related to (F \\U) ∪ V . Figure 30 shows the move F → (F \\U) ∪ V within each prism Pi. Whereas the disk U intersects C in some number of saddle disks and no crossing bands, the disk V intersects C in no saddle disks and some number of crossing bands. Moreover, (11) implies that either U ∩ C 6= ∅ or V contains a crossing band. Therefore, the move F → (F \\U) ∪ V decreases the complexity of F : ||(F \\U) ∪ V || < ||F ||. Since F was already isotoped to minimize ||F ||, this move is actually a re-plumbing move, and V was not a fake plumbing cap.

35A caveat analogous to the one in Note 34 applies here as well. 36 THOMAS KINDRED

After performing the re-plumbing move, isotope the resulting sur- face into good position so as to minimize its complexity. If it still intersects H+, then perform another re-plumbing move so as further to decrease its complexity. Repeat in this manner, performing iso- topy and re-plumbing moves until the resulting surface is disjoint from H+ and thus is isotopic to B.  One can visualize the re-plumbing move in the proof of Theorem 5.26 in the following alternative way. Reflect the disk X0 across the projection sphere. The resulting surface has arcs of self-intersection, but each of these arcs has one endpoint on ∂νL and the other in its interior, on an arc of γ0 ∩ W . “Unzip” the surface along these arcs. Flyping Theorem. Any two reduced alternating diagrams of the same prime, nonsplit link are flype-related. Proof. Let D,D0 be reduced alternating diagrams of a prime nonsplit link L with respective chessboards B,W and B0,W 0, where B,B0 are positive-definite. Then B and B0 are plumb-related, by Theorem 5.26, as are W and W 0. Therefore, D and D0 are flype-related, by Theorem 4.11.  Thank you to Colin Adams for posing a question about flypes and chessboards during SMALL 2005 which eventually led to Fig- ure 13. Thank you to Josh Howie and Alex Zupan for suggesting improvements to the exposition of this paper. Thank you to Josh Greene for suggesting such improvements and for encouraging me to think about this problem when he hosted me at Boston College a few years ago.

References [1] C. Adams, T. Kindred, A classification of spanning surfaces for alternating links, Alg. Geom. Topol. 13 (2013), no. 5, 2967-3007. [2] C. McA. Gordon, R.A. Litherland, On the signature of a link, Invent. Math. 47 (1978), no. 1, 53-69. [3] J. Greene, Alternating links and definite surfaces, with an appendix by A. Juhasz, M Lackenby, Duke Math. J. 166 (2017), no. 11, 2133-2151. [4] J. Howie, A characterisation of alternating knot exteriors, Geom. Topol. 21 (2017), no. 4, 2353-2371. [5] T. Kindred, Alternating links have representativity 2, Alg. Geom. Topol. 18 (2018), no. 6, 3339-3362. [6] L.H. Kauffman, State models and the Jones polynomial, Topology 26 (1987), no. 3, 395-407. [7] W. Menasco, Closed incompressible surfaces in alternating knot and link complements, Topology 23 (1984), no. 1, 37-44. [8] W. Menasco, M. Thistlethwaite, The Tait flyping conjecture, Bull. Amer. Math. Soc. (N.S.) 25 (1991), no. 2, 403-412. [9] W. Menasco, M. Thistlethwaite, The classification of alternating links, Ann. of Math. (2) 138 (1993), no. 1, 113-171. A GEOMETRIC PROOF OF THE FLYPING THEOREM 37

[10] K. Murasugi, Jones polynomials and classical conjectures in , Topology 26 (1987), no. 2, 187-194. [11] P.G. Tait, On Knots I, II, and III, Scientific papers 1 (1898), 273-347. [12] M.B. Thistlethwaite, A spanning tree expansion of the Jones polynomial, Topology 26 (1987), no. 3, 297-309. [13] M.B. Thistlethwaite, On the Kauffman polynomial of an adequate link, In- vent. Math. 93 (1988), no. 2, 285-296. [14] M.B. Thistlethwaite, On the algebraic part of an alternating link, Pacific J. Math. 151 (1991), no. 2, 317-333. [15] V.G. Turaev, A simple proof of the Murasugi and Kauffman theorems on alternating links, Enseign. Math. (2) 33 (1987), no. 3-4, 203-225.

Department of Mathematics, University of Nebraska, Lincoln, Ne- braska 68588-0130, USA