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

Integer surgeries rational homology cobordant to lens spaces

Antony T.H. Fung
Abstract.

The Cyclic Surgery Theorem and Moser’s work on surgeries on torus knots imply that for any non-trivial knot in S3S^{3}, there are at most two integer surgeries that produce a lens space. This paper investigates how many positive integer surgeries on a given knot in S3S^{3} can produce a manifold rational homology cobordant to a lens space. Tools include Greene and McCoy’s work on changemaker lattices which come from Heegaard Floer dd-invariants, and Aceto-Celoria-Park’s work on rational cobordisms and integral homology which is based on Lisca’s work on lens spaces.

1. Introduction

Lens spaces are one of the simplest classes of 3-manifolds, and Dehn surgery is one of the simplest ways of constructing 3-manifolds. There are a lot of famous theorems and conjectures that concern when surgery on a knot in S3S^{3} produces a lens space. For example, the Berge conjecture concerns “which knots”, the Cyclic Surgery Theorem [7] concerns “which slopes”, and Greene’s work on the lens space realization problem [9] answers the question of “which lens spaces”.

This paper addresses a generalization of these questions, namely when surgery on a knot in S3S^{3} is smoothly rational homology cobordant to a lens space. The following definition is central to this paper:

Definition 1.1.

A lensbordant surgery on a knot in an integer homology 3-sphere is a positive integer surgery which is smoothly rational homology cobordant to a lens space.

We make the following conjecture, which is an analog of the Cyclic Surgery Theorem.

Conjecture 1.2.

There exists a positive integer NN such that for all knots KK in S3S^{3} satisfying ν+(K)0\nu^{+}(K)\neq 0, there are less than NN lensbordant surgeries on KK.

The non-negative integer ν+(K)\nu^{+}(K) is a lower bound on the 4-ball genus g4(K)g_{4}(K) coming from work by Rasmussen [15] and Ni and Wu [14]. When KK is an LL-space knot, ν+(K)\nu^{+}(K) coincides with the genus g(K)g(K).

We are not aware of any knot KK in S3S^{3} satisfying ν+(K)0\nu^{+}(K)\neq 0 having more than 5 lensbordant surgeries, though we also do not have a good reason to believe that such a knot cannot exist. There are examples of knots with 5 known lensbordant surgeries: When KK is the torus knot Tp,p+1T_{p,p+1}, the (p2+p+1)(p^{2}+p+1)-surgery and the (p2+p1)(p^{2}+p-1)-surgery are both lens spaces, while the p2p^{2}-surgery and the (p+1)2(p+1)^{2}-surgery both bound a smooth rational ball [2, Th. 1.4], and hence being smoothly rational homology cobordant to L(4,3)L(4,3). This gives 4 lensbordant surgeries. In particular, when p=2p=2, the 8-surgery is smoothly rational homology cobordant to L(2,1)L(2,1) [3, Th. 1.3]. When p=3p=3, the 12-surgery produces L(4,1)#L(3,2)L(4,1)\#L(3,2), which is smoothly rational homology coboardant to L(3,2)L(3,2). When p=4p=4, the 20-surgery produces L(5,1)#L(4,3)L(5,1)\#L(4,3), which is smoothly rational homology coboardant to L(5,1)L(5,1). Hence, when p=2,3,4p=2,3,4, there are 5 known lensbordant surgeries. Note that in this paper, we take the convention where L(p,q)L(p,q) is the pq-\tfrac{p}{q}-surgery on the unknot in S3S^{3}.

The first main result of this paper is

Theorem 1.3.

Let KK be a knot in S3S^{3}. If ν+(K)0\nu^{+}(K)\neq 0, there is a finite number of lensbordant surgeries on KK.

This is implied by Theorem 1.7 and Corollary 1.9, which are stated later in this section and proved in Section 4 and Section 5 respectively.

We do not know whether the condition ν+(K)0\nu^{+}(K)\neq 0 can be weakened. However, we know that it is necessary to have some condition because the conclusion of the theorem is false whenever KK is a slice knot. When KK is a slice knot, all positive integer surgeries are integer homology cobordant to a lens space. We do not know whether there is any non-slice knot with an infinite number of lensbordant surgeries.

As a complement to Theorem 1.3, we have the following result about knots with vanishing ν+\nu^{+}:

Theorem 1.4.

Let KK be a knot in S3S^{3} with ν+(K)=0\nu^{+}(K)=0 and let mm be a positive integer. If the mm-surgery on KK is lensbordant, then it is smoothly rational homology cobordant to L(m,m1)L(m,m-1).

Remark: When m=1m=1, “L(m,m1)L(m,m-1)” is referring to S3S^{3}.

Theorem 1.4 is proved in Section 4.

Aceto, Celoria, and Park defined the notion of a reduced lens space [1, Def. 2.2]. Any lens space is smoothly rational homology cobordant to a reduced lens space. For simplicity, in this paper we say that S3S^{3} is reduced with p=1p=1, so when we say “reduced lens space” it also includes S3S^{3}.

Theorem 1.5.

If a positive integer surgery on a knot in S3S^{3} is smoothly rational homology cobordant to a reduced lens space, then that reduced lens space must also be a positive integer surgery on a knot in S3S^{3}.

Section 3 of this paper is dedicated to proving Theorem 1.5.

Questions: Is there any interesting relationship between the two knots mentioned in Theorem 1.5? Is there any nice construction for knots in S3S^{3} with lensbordant surgeries similar to what Berge [4] did for knots in S3S^{3} with lens space surgeries?

If a positive integer surgery on a knot KK in an integer homology 3-sphere is smoothly rational cobordant to a reduced L(p,q)L(p,q), the surgery slope must be r2pr^{2}p for some positive integer rr [1, Th 1.1]. This motivates the following definition:

Definition 1.6.

Let r,pr,p be positive integers. A surgery on a knot in an integer homology 3-sphere is (r,p)(r,p)-lensbordant if the surgery slope is r2pr^{2}p and the resulting manifold is smoothly rational homology cobordant to some reduced L(p,q)L(p,q).

Our next result is an analog of the Cyclic Surgery Theorem in the direction of Conjecture 1.2. The result states that for most knots KK, when rr is fixed, the number of possible surgery slopes is very limited. More precisely:

Theorem 1.7.

Let KK be a knot in S3S^{3} with ν+(K)0\nu^{+}(K)\neq 0. Then,
(1) When r=1r=1, there are at most 3 (r,p)(r,p)-lensbordant surgeries on KK.
(2) When rr is even, there are at most 2 (r,p)(r,p)-lensbordant surgeries on KK.
(3) When r3r\geq 3 is odd, there is at most 1 (r,p)(r,p)-lensbordant surgery on KK.

Section 4 of this paper is dedicated to proving Theorem 1.7.

The proofs of Theorem 1.5 and Theorem 1.7 involve the changemaker vector σ\sigma described in [9]. Every lens space L(p,q)L(p,q) is the boundary of a canonical negative definite plumbing 4-manifold X(p,q)X(p,q) that arises from the Hirzebruch-Jung continued fraction of pq\tfrac{p}{q}. Let nn be the rank of H2(X(p,q);)H_{2}(X(p,q);\mathbb{Z}). When L(p,q)L(p,q) is a positive integer surgery on a knot in S3S^{3}, the intersection form of X(p,q)X(p,q) embeds in n+1\mathbb{Z}^{n+1} as the orthogonal complement of some vector σ\sigma with entries that satisfy certain combinatorial properties [10]. We say that σ\sigma is a changemaker associated with L(p,q)L(p,q). Note that L(p,q)L(p,q) can have multiple associated changemakers. In this paper, whenever we say that a lensbordant surgery is associated with a changemaker σ\sigma, we mean that the surgery is rational homology cobordant to a reduced L(p,q)L(p,q) associated with σ\sigma. Because of Theorem 1.5, every lensbordant surgery is associated with some changemaker σ\sigma.

In Section 5, we establish some bounds for the surgery slope r2pr^{2}p. At the start of Section 5, we show that

Theorem 1.8.

Suppose there is an (r,p)(r,p)-lensbordant surgery on a knot KK in S3S^{3} with associated changemaker σ\sigma, then

2ν+(K)+2(r1)r2pr|σ|12ν+(K).2\nu^{+}(K)+2(r-1)\geq r^{2}p-r|\sigma|_{1}\geq 2\nu^{+}(K).

Using the fact that |σ|1p|\sigma|_{1}\leq p, this implies

Corollary 1.9.

Suppose there is an (r,p)(r,p)-lensbordant surgery on a knot KK in S3S^{3} with associated changemaker σ\sigma, then
(1) If r3r\geq 3, then 2r2(r1)(r2)ν+(K)r2p2ν+(K)+r\dfrac{2r^{2}}{(r-1)(r-2)}\nu^{+}(K)\geq r^{2}p\geq 2\nu^{+}(K)+r
(2) If r2r\geq 2, then 4ν+(K)+5r2p2ν+(K)+r4\nu^{+}(K)+5\geq r^{2}p\geq 2\nu^{+}(K)+r

Theorem 1.8 and Corollary 1.9 are proved in Section 5.

For every knot KK in S3S^{3}, based on work by Rasmussen [15] and Ni and Wu [14], there is an associated sequence of non-negative integers Vi(K)V_{i}(K). The non-negative integer ν+(K)\nu^{+}(K) is the smallest index ii such that Vi(K)=0V_{i}(K)=0. Also, when KK is an LL-space knot, the coefficients Vi(K)V_{i}(K) coincide with the torsion coefficients ti(K)t_{i}(K) which can be computed from the Alexander polynomial.

In Section 6, we will look in depth into the combinatorial conditions established in Section 4. We will show that

Theorem 1.10.

Suppose there is an (r,p)(r,p)-lensbordant surgery on a knot KK in S3S^{3}. If r,pr,p are both even, the surgery slope r2pr^{2}p must equal 8V0(K)8V_{0}(K).

Note that when r,pr,p are not both even, the surgery slope r2pr^{2}p is still very close to 8V0(K)8V_{0}(K). See Theorem 6.1 in Section 6 for a precise statement.

Because of Theorem 1.5, we can focus on lens spaces that are positive integer surgeries on knots in S3S^{3}. These lens spaces are classified. Berge [4] found examples of integer surgeries on knots in S3S^{3} producing lens spaces, and Rasmussen [16] organized them nicely into a list of Berge types. Greene [9] showed that this is the complete list of lens spaces that are integer surgeries on knots in S3S^{3}. He showed this by analyzing the possible structures of changemakers that arise from positive integer surgeries on knots in S3S^{3} producing lens spaces. Greene concluded that such a changemaker must be in one of the 33 forms that he wrote down in [9], and then noted that for all 33 such structures, the resulting lens spaces are within this list of Berge types. Among these 33 structures in [9], 26 are classified as “small families” and 7 are classified as “large families”.

One way to tackle Conjecture 1.2 is to show that for each of these 33 changemaker structures described by Greene, there exists some NN such that for all knots KK in S3S^{3}, there are at most NN lensbordant surgeries associated with a changemaker having that structure. In Section 7, we do this for one of the small families. This small family structure has changemakers in the form

σ=(4s+3,2s+1,s+1,s,1,,1s copies of 1’s)\sigma=\begin{pmatrix}4s+3,&2s+1,&s+1,&s,&\smash{\underbrace{\begin{matrix}1,&\dots&,&1\end{matrix}}_{s\text{ copies of }1\text{'s}}}\end{pmatrix}

where s2s\geq 2.

In this paper, we show that

Theorem 1.11.

Any knot in S3S^{3} has at most 7 lensbordant surgeries with an associated changemaker coming from family mentioned above.

Theorem 1.11 is proved in Section 7.

We believe that similar techniques can be use on all 26 small families. However, we do not see how this approach extends to the 7 large families. Some large families are expected to be difficult to tackle because they contain infinite families of non-reduced lens spaces, which are being studied by the author in a separate project.

So far, everything we have discussed addresses surgeries on knots in S3S^{3}. One may ask whether this generalizes to knots in other integer homology 3-spheres. Caudell addressed the lens space realization problem for knots in the Poincaré homology sphere [6] 𝒫\mathcal{P}. In 𝒫\mathcal{P}, we need to make more assumptions. We will show that

Theorem 1.12.

Suppose r2pr^{2}p-surgery on a knot K𝒫K\subset\mathcal{P} produces an L-space that is smoothly 2\mathbb{Z}_{2}-homology cobordant to a reduced L(p,q)L(p,q). Let

i(K,r,p)={2g(K)+r if p is odd2g(K) if p is eveni(K,r,p)=\begin{cases}2g(K)+r\text{ if }p\text{ is odd}\\ 2g(K)\quad\ \ \text{ if }p\text{ is even}\end{cases}

If r2pi(K,r,p)r^{2}p\geq i(K,r,p), then L(p,q)L(p,q) must be a positive integer surgery on a knot in 𝒫\mathcal{P}.
If r2p<i(K,r,p)r^{2}p<i(K,r,p), then L(p,q)L(p,q) must be a positive integer surgery on a knot in S3S^{3}.

Theorem 1.12 is proved in Section 8. Note that we do not know whether the conditions in Theorem 1.12 can be weakened.

Question: Does Theorem 1.12 generalize to knots in other integer homology spheres? One potentially interesting candidate to consider would be Σ(2,3,7)\Sigma(2,3,7). In [19], Tange gave some examples of positive integer surgeries on knots in Σ(2,3,7)\Sigma(2,3,7) yielding lens spaces.

Convention: In this paper, unless specified, homology and cohomology always take integer coefficients.

Acknowledgements: The author would like to thank his adviser Joshua Greene for inspiring this project and also for many helpful conversations throughout the project. The author would also like to thank Duncan McCoy for pointing out that using VV coefficients instead of torsion coefficients allows results to be stated for general knots in S3S^{3} instead of just L-space knots in S3S^{3}, and thank Paolo Aceto for mentioning examples of knots in S3S^{3} with 5 lensbordant surgeries.

2. Analyzing the Aceto-Celoria-Park map

In [1], Aceto, Celoria, and Park showed that when a closed 3-manifold YY is smoothly rational homology cobordant to a reduced connected sum of lens spaces LYL_{Y} (see [1, Def. 2.2]), there is an injective map H1(LY)H1(Y)H_{1}(L_{Y})\rightarrow H_{1}(Y). They proved this by showing that H1(LY)H_{1}(L_{Y}) injects into a group that H1(Y)H_{1}(Y) surjects onto, and note that if GGG\rightarrow G^{\prime} is a surjective map between finite abelian groups then there exists an injection GGG^{\prime}\rightarrow G.

The goal of this section is to show that this interacts well with the spinc\text{spin}^{\text{c}} structures (and hence the dd-invariants).

We use the notations in [1]. Let WW be the smooth rational cobordism between LYL_{Y} and YY. Let PP be the canonical negative definite plumbed 4-manifold bounded by LYL_{Y}. Let X:=PLWX:=P\cup_{L}W.

Aceto, Celoria, and Park showed that the intersection forms of PP and XX are isomorphic [1, Prop. 4.1], and its proof implies that the isomorphism is induced by the inclusion PXP\rightarrow X. Since H2(P)H_{2}(P) is free, we have the decomposition

H2(X)=Free(H2(X))Torsion(H2(X))H_{2}(X)=Free(H_{2}(X))\oplus Torsion(H_{2}(X))

with Free(H2(X))Free(H_{2}(X)) being the image of H2(P)iH2(X)H_{2}(P)\xrightarrow{i_{*}}H_{2}(X).

Lemma 2.1.

The map H1(LY)iH1(W)H_{1}(L_{Y})\xrightarrow{i_{*}}H_{1}(W) induced by inclusion is injective.

Proof.

By excision and the neighborhood collar theorem, the inclusion (P,LY)(X,W)(P,L_{Y})\rightarrow(X,W) induces isomorphisms H(P,LY)H(X,W)H_{*}(P,L_{Y})\cong H_{*}(X,W).

Consider the long exact seqence of the pair (X,W)(X,W):

H2(W)iH2(X)H2(X,W)\rightarrow H_{2}(W)\xrightarrow{i_{*}}H_{2}(X)\rightarrow H_{2}(X,W)\rightarrow

Since H1(P)=0H_{1}(P)=0, H2(X,W)H2(P,LY)H2(P)H_{2}(X,W)\cong H_{2}(P,L_{Y})\cong H^{2}(P) is free, and hence the image of the map H2(W)iH2(X)H_{2}(W)\xrightarrow{i_{*}}H_{2}(X) contains all torsion elements of H2(X)H_{2}(X). Now we consider the Mayer-Vietoris sequence

0H2(P)H2(W)iiH2(X)H1(LY)iH1(W)0\rightarrow H_{2}(P)\oplus H_{2}(W)\xrightarrow{i_{*}-i_{*}}H_{2}(X)\rightarrow H_{1}(L_{Y})\xrightarrow{i_{*}}H_{1}(W)\rightarrow

Since the image of the map H2(P)iH2(X)H_{2}(P)\xrightarrow{i_{*}}H_{2}(X) is Free(H2(X))Free(H_{2}(X)) and the image of the map H2(W)iH2(X)H_{2}(W)\xrightarrow{i_{*}}H_{2}(X) contains all torsion elements, we know that the map H2(P)H2(W)iiH2(X)H_{2}(P)\oplus H_{2}(W)\xrightarrow{i_{*}-i_{*}}H_{2}(X) is surjective, and hence the map H1(LY)iH1(W)H_{1}(L_{Y})\xrightarrow{i_{*}}H_{1}(W) is injective. ∎

Consider the diagram


0{0}H2(X){H_{2}(X)}H2(X,Y){H_{2}(X,Y)}H1(Y){H_{1}(Y)}  H2(X,W){H_{2}(X,W)}0{0}H2(P){H_{2}(P)}H2(P,LY){H_{2}(P,L_{Y})}H1(LY){H_{1}(L_{Y})}0{0}ϕ\scriptstyle{\phi}b\scriptstyle{b}h\scriptstyle{h}i\scriptstyle{i_{*}}\scriptstyle{\cong}

The two rows are the long exact sequences of the pairs (X,Y)(X,Y) and (P,LY)(P,L_{Y}). ii_{*} is induced by the inclusion PXP\rightarrow X. The map H2(X,Y)H2(X,W)H_{2}(X,Y)\rightarrow H_{2}(X,W) comes from the long exact sequence of the triple (Y,W,X)(Y,W,X). The map H2(P,LY)H2(X,W)H_{2}(P,L_{Y})\rightarrow H_{2}(X,W) is excision.

For notational convenience in later parts of this section, we label the map H2(X)H2(X,Y)H_{2}(X)\rightarrow H_{2}(X,Y) as ϕ\phi, the map H2(P)H2(P,LY)H_{2}(P)\rightarrow H_{2}(P,L_{Y}) as hh, and the map H2(X,Y)H1(Y)H_{2}(X,Y)\rightarrow H_{1}(Y) as bb.

Lemma 2.2.

This diagram commutes.

Proof.

Given a cycle αC2(P)\alpha\in C_{2}(P), both directions map α\alpha to [α]C2(X)C2(W)[\alpha]\in\dfrac{C_{2}(X)}{C_{2}(W)}, where we view αC2(P)C2(X)\alpha\in C_{2}(P)\subset C_{2}(X).

As mentioned in the proof of Lemma 2.1, H2(X,W)H_{2}(X,W) is free. Hence, in the map H2(X,Y)H2(X,W)H_{2}(X,Y)\rightarrow H_{2}(X,W), all torsion elements of H2(X,Y)H_{2}(X,Y) are mapped to 0. Hence, it induces a map

f:H2(X,Y)Torsion(H2(X,Y))H2(X,W)f:\dfrac{H_{2}(X,Y)}{Torsion(H_{2}(X,Y))}\rightarrow H_{2}(X,W)

Treating Free(H2(X))Free(H2(X)){0}Free(H_{2}(X))\cong Free(H_{2}(X))\oplus\{0\} as a subgroup of H2(X)H_{2}(X), we obtain the following new commutative diagram where everything except H1(LY)H_{1}(L_{Y}) is free.


0{0}Free(H2(X)){Free(H_{2}(X))}H2(X,Y)Torsion(H2(X,Y)){\dfrac{H_{2}(X,Y)}{Torsion(H_{2}(X,Y))}}  H2(X,W){H_{2}(X,W)}0{0}H2(P){H_{2}(P)}H2(P,LY){H_{2}(P,L_{Y})}H1(LY){H_{1}(L_{Y})}0{0}g\scriptstyle{g}f\scriptstyle{f}h\scriptstyle{h}\scriptstyle{\cong}\scriptstyle{\cong}

The map on top, which we labeled as gg, is obtained by composing ϕ|Free(H2(X))\phi|_{Free(H_{2}(X))} with the projection

H2(X,Y)H2(X,Y)Torsion(H2(X,Y))H_{2}(X,Y)\rightarrow\dfrac{H_{2}(X,Y)}{Torsion(H_{2}(X,Y))}

Using notations in [1, §4], let QXQ_{X} be the matrix representation of the intersection form of XX, which also represents the map gg under some suitable basis. Similarly, let QPQ_{P} be the matrix representation of the intersection form of PP, which also represents the map H2(P)H2(P,LY)H_{2}(P)\rightarrow H_{2}(P,L_{Y}) under some suitable basis.

Since (H2(P),QP)(H_{2}(P),Q_{P}) and (H2(X),QX)(H_{2}(X),Q_{X}) are isomorphic [1, Prop. 4.1], we know that det(QP)=det(QX)det(Q_{P})=det(Q_{X}). Since |H1(LY)||H_{1}(L_{Y})| is finite, we know that det(QP)0det(Q_{P})\neq 0. Therefore, ff has has determinant 1, and hence we know that ff is an isomorphism.

By the commutativity of the diagram, we know that the image of the map fgf\circ g coincides with the image of the map H2(P)H2(P,LY)H2(X,W)H_{2}(P)\rightarrow H_{2}(P,L_{Y})\cong H_{2}(X,W). Hence, ff induces an isomorphism

f¯:H2(X,Y)GH2(P,LY)h(H2(P))\bar{f}:\dfrac{H_{2}(X,Y)}{G}\rightarrow\dfrac{H_{2}(P,L_{Y})}{h(H_{2}(P))}

where GG is the subgroup of H2(X,Y)H_{2}(X,Y) generated by elements in ϕ(Free(H2(X)))\phi(Free(H_{2}(X))) and Torsion(H2(X,Y))Torsion(H_{2}(X,Y)).

A homomorphism cannot map a torsion element to a non-torsion element. Therefore, the image of Torsion(H2(X))Torsion(H_{2}(X)) under ϕ\phi lies within Torsion(H2(X,Y))Torsion(H_{2}(X,Y)). Hence, GG is exactly the subgroup of H2(X,Y)H_{2}(X,Y) generated by elements in ϕ(H2(X))\phi(H_{2}(X)) and Torsion(H2(X,Y))Torsion(H_{2}(X,Y)).

We go back to our first diagram


0{0}H2(X){H_{2}(X)}H2(X,Y){H_{2}(X,Y)}H1(Y){H_{1}(Y)}  H2(X,W){H_{2}(X,W)}0{0}H2(P){H_{2}(P)}H2(P,LY){H_{2}(P,L_{Y})}H1(LY){H_{1}(L_{Y})}0{0}ϕ\scriptstyle{\phi}b\scriptstyle{b}h\scriptstyle{h}i\scriptstyle{i_{*}}\scriptstyle{\cong}

Let H1(Y)¯\overline{H_{1}(Y)} be the image of bb, which is a subgroup of H1(Y)H_{1}(Y).

We see that H1(LY)H2(P,LY)h(H2(P))H_{1}(L_{Y})\cong\dfrac{H_{2}(P,L_{Y})}{h(H_{2}(P))} and H2(X,Y)ϕ(H2(X))H1(Y)¯\dfrac{H_{2}(X,Y)}{\phi(H_{2}(X))}\cong\overline{H_{1}(Y)}. Hence, we can view f¯\bar{f} as an isomorphism

H1(Y)¯b(Torsion(H2(X,Y)))H1(LY)\dfrac{\overline{H_{1}(Y)}}{b(Torsion(H_{2}(X,Y)))}\rightarrow H_{1}(L_{Y})

Now we compose this with the projection

H1(Y)¯H1(Y)¯b(Torsion(H2(X,Y)))\overline{H_{1}(Y)}\rightarrow\dfrac{\overline{H_{1}(Y)}}{b(Torsion(H_{2}(X,Y)))}

We obtain a surjective map H1(Y)¯H1(LY)\overline{H_{1}(Y)}\rightarrow H_{1}(L_{Y}) which we denote as ss.

Now, we consider the maps H1(LY)H1(W)H_{1}(L_{Y})\rightarrow H_{1}(W) and H1(Y)H1(W)H_{1}(Y)\rightarrow H_{1}(W) induced by inclusion.

Lemma 2.3.

The diagram H1(Y)¯{\overline{H_{1}(Y)}}H1(W){H_{1}(W)}H1(LY){H_{1}(L_{Y})}i|H1(Y)¯\scriptstyle{i_{*}|_{\overline{H_{1}(Y)}}}s\scriptstyle{s}i\scriptstyle{i_{*}} commutes

Proof.

Let yH1(Y)¯y\in\overline{H_{1}(Y)}.

Let αC2(X)\alpha\in C_{2}(X) be such that α\alpha represents a cycle in C2(X,Y)C_{2}(X,Y) that represents an element in H2(X,Y)H_{2}(X,Y) that gets mapped to yy under bb. By considering how the map bb works in the long exact sequence of a pair, we see that yy is represented by αC1(Y)C1(X)\partial\alpha\in C_{1}(Y)\subset C_{1}(X).

By considering the chain maps being used in defining the Mayer-Vietoris sequence for PW=XP\cup W=X, we can write α=βγ\alpha=\beta-\gamma, where βC2(P)\beta\in C_{2}(P) and γC2(W)\gamma\in C_{2}(W). Under

H2(X,Y)H2(X,W)H2(P,LY)H1(LY)H_{2}(X,Y)\rightarrow H_{2}(X,W)\cong H_{2}(P,L_{Y})\rightarrow H_{1}(L_{Y})

The element [α]H2(X,Y)[\alpha]\in H_{2}(X,Y) is being mapped to [β]H2(P,LY)[\beta]\in H_{2}(P,L_{Y}), which is then being mapped to [β]H1(LY)[\partial\beta]\in H_{1}(L_{Y}), where βC1(LY)C1(P)\partial\beta\in C_{1}(L_{Y})\subset C_{1}(P).

Now we have s(y)=[β]s(y)=[\partial\beta]. From α=βγ\alpha=\beta-\gamma, we know that γ=βα\partial\gamma=\partial\beta-\partial\alpha. Hence, inside C1(W)C_{1}(W), β\partial\beta and α\partial\alpha differ by a boundary element, and therefore represent the same element in H1(W)H_{1}(W). ∎

Lemma 2.4.

Let A1,,AkA_{1},\dots,A_{k} be distinct spinc\text{spin}^{\text{c}} structures on LYL_{Y}. Then there exists distinct spinc\text{spin}^{\text{c}} structures B1,,BkB_{1},\dots,B_{k} on YY such that for each ii, AiA_{i} and BiB_{i} are restrictions of the same spinc\text{spin}^{\text{c}} structure on WW.

Proof.

Consider the long exact sequence

  H2(W){H^{2}(W)}H2(LY)H2(Y){H^{2}(L_{Y})\oplus H^{2}(Y)}H3(W,LYY){H^{3}(W,L_{Y}\cup Y)}  H1(LY)H1(Y){H_{1}(L_{Y})\oplus H_{1}(Y)}H1(W){H_{1}(W)}i\scriptstyle{i^{*}}i\scriptstyle{i_{*}}\scriptstyle{\cong}\scriptstyle{\cong}

Let 𝔰\mathfrak{s} be a spinc\text{spin}^{\text{c}} structure on WW. For any ll, Al=al𝔰|LYA_{l}=a_{l}\cdot\mathfrak{s}|_{L_{Y}} for some alH2(LY)H1(LY)a_{l}\in H^{2}(L_{Y})\cong H_{1}(L_{Y}). By Lemma 2.3 and the surjectivity of ss, we know that there exists some blH1(Y)¯H1(Y)H2(Y)b_{l}\in\overline{H_{1}(Y)}\subset H_{1}(Y)\cong H^{2}(Y) such that (al,bl)(a_{l},b_{l}) is in the kernal of ii_{*}, and hence it is in the image of ii^{*}. Let clH2(W)c_{l}\in H^{2}(W) be such that i(cl)=(al,bl)i^{*}(c_{l})=(a_{l},b_{l}). We set Bl:=bl𝔰|YB_{l}:=b_{l}\cdot\mathfrak{s}|_{Y}. Then cl𝔰c_{l}\cdot\mathfrak{s} restricts to AlA_{l} and BlB_{l} respectively. The BlB_{l}’s are distinct because given any blH1(Y)¯b_{l}\in\overline{H_{1}(Y)}, we can use ss to find the corresponding ala_{l} in H1(LY)H_{1}(L_{Y}). ∎

Corollary 2.5.

The list of dd-invariants of YY is extended from the list of dd-invariants of LYL_{Y}.

Proof.

That is because dd-invariants are invariant under smooth rational homology cobordism. ∎

3. Proof of Theorem 1.5

This section is dedicated to proving Theorem 1.5.

Theorem 1.5.

If a positive integer surgery on a knot in S3S^{3} is smoothly rational homology cobordant to a reduced lens space, then that reduced lens space must also be a positive integer surgery on a knot in S3S^{3}.

We continue the work in Section 2, using the same notations. But now LYL_{Y} is a reduced L(p,q)L(p,q), and YY is the r2pr^{2}p-surgery on KK. We denote the r2pr^{2}p-surgery on KK as Kr2pK_{r^{2}p}.

The goal of this section is to show that L(p,q)L(p,q) is a positive integer surgery on a knot in S3S^{3}. To show that, we will use Greene’s work on the lens space realization problem [9, Th. 1.7]. To state the statement of [9, Th. 1.7], we need the following definition:

Definition 3.1.

[9, Def. 1.5] A changemaker vector is a vector σ=(σ0,,σn)>0n+1\sigma=(\sigma_{0},\dots,\sigma_{n})\in\mathbb{Z}^{n+1}_{>0} such that for all kk\in\mathbb{Z} satisfying 0kσ0++σn0\leq k\leq\sigma_{0}+\dots+\sigma_{n}, there exists some A{σ0,,σn}A\subseteq\{\sigma_{0},\dots,\sigma_{n}\} such that the sum of the numbers in AA equals to kk.

Note that we requires the σ\sigma to be in >0n+1\mathbb{Z}^{n+1}_{>0}, which implies the entries of σ\sigma to be all non-zero. This is consistent with [9], but different from some other papers (for example [10]).

In this paper, we will choose a basis in a way such that σ0σn\sigma_{0}\geq\dots\geq\sigma_{n}, which is the opposite of the ordering used in [9, Def. 1.5].

Let qq^{\prime} be such that qq1qq^{{}^{\prime}}\equiv 1 (mod pp). Let Λ(p,q)\Lambda(p,q) be the lattice coming from the intersection pairing on PP. Recall from Section 2 that PP is the canonical negative definite plumbing bounded by LYL_{Y}, and in this section we are setting LY=L(p,q)L_{Y}=L(p,q).

Greene showed that Λ(p,q)\Lambda(p,q) embeds as the orthogonal complement to a changemaker vector in n+1\mathbb{Z}^{n+1} if and only if at least one of L(p,q)L(p,q), L(p,q)L(p,q^{{}^{\prime}}) appears on Berge’s list of lens spaces obtained from a positive integer surgery on a knot in S3S^{3} [9, Th. 1.7].

Since L(p,q)L(p,q^{{}^{\prime}}) is homeomorphic to L(p,q)L(p,q) in an orientation preserving manner, to prove Theorem 1.5, it suffices to show that Λ(p,q)\Lambda(p,q) embeds as the orthogonal complement to a changemaker vector in n+1\mathbb{Z}^{n+1}. This is what we will do in this section.

Similar to the set up in [10, §2], we let Wr2pW_{r^{2}p} be the 4-manifold obtained by attaching a r2pr^{2}p-framed 2-handle along KS3=D4K\subset S^{3}=\partial D^{4}. By construction, we have H1(Wr2p)=0H_{1}(W_{r^{2}p})=0. The manifold Wr2pW_{r^{2}p} has boundary Kr2pK_{r^{2}p}. Let ZZ be the closed 4-manifold XKr2pWr2pX\cup_{K_{r^{2}p}}W_{-r^{2}p}. H2(Wr2p)H_{2}(-W_{r^{2}p}) is generated by the class of a surface Σ\Sigma obtained by gluing the core of the 2-handle to a Seifert surface of KK. Σ\Sigma has the property that the intersection pairing of [Σ][\Sigma] with itself gives r2p-r^{2}p. Each spinc\text{spin}^{\text{c}} structure on Kr2pK_{r^{2}p} comes with a label i{0,,r2p1}i\in\{0,\dots,r^{2}p-1\}. This label has the property that for all 𝔰Spinc(Wr2p)\mathfrak{s}\in\text{Spin}^{\text{c}}(-W_{r^{2}p}) that extends the spinc\text{spin}^{\text{c}} structure on Kr2pK_{r^{2}p} with label ii, we have c1(𝔰),[Σ]+r2p2i\langle c_{1}(\mathfrak{s}),[\Sigma]\rangle+r^{2}p\equiv 2i (mod 2r2p2r^{2}p).

Note that we do not completely follow the notations in [10, §2]. We defined WW to be the cobordism in order to be consistent with [1], and later in this chapter we will define σ\sigma slightly differently from [10, §2].


[Uncaptioned image]

Figure 1. The set up of Section 3.


Lemma 3.2.

H2(P)H_{2}(P) embeds primitively in H2(Z)H_{2}(Z), in the sense that for any kk\in\mathbb{Z} and zH2(Z)z\in H_{2}(Z), if kzH2(P)kz\in H_{2}(P), then zz must equals to p+τp+\tau for some pH2(P)p\in H_{2}(P) and τ\tau being a torsion element in H2(Z)H_{2}(Z).

Proof.

Suppose kzkz is in H2(P)H_{2}(P).

Consider the Mayer-Vietoris sequence

0H2(P)H2(W)H2(PW)H1(L(p,q))H1(P)H1(W)0\rightarrow H_{2}(P)\oplus H_{2}(W)\rightarrow H_{2}(P\cup W)\rightarrow H_{1}(L(p,q))\rightarrow H_{1}(P)\oplus H_{1}(W)\rightarrow\dots

By Lemma 2.1, we know that the map H2(P)H2(W)H2(PW)H_{2}(P)\oplus H_{2}(W)\rightarrow H_{2}(P\cup W) is an isomorphism.

Since H2(P)H_{2}(P) embeds in H2(PW)H_{2}(P\cup W), we can also treat kzkz as an element of H2(PW)H_{2}(P\cup W). Now, consider the exact sequence of a pair

H3(Z,PW)H2(PW)H2(Z)H2(Z,PW)\dots\rightarrow H_{3}(Z,P\cup W)\rightarrow H_{2}(P\cup W)\rightarrow H_{2}(Z)\rightarrow H_{2}(Z,P\cup W)\rightarrow\dots

By excision, we have

H3(Z,PW)H3(Wr2p,Kr2p)H1(Wr2p)0H_{3}(Z,P\cup W)\cong H_{3}(-W_{r^{2}p},K_{r^{2}p})\cong H^{1}(-W_{r^{2}p})\cong 0

Also by excsion, we have

H2(Z,PW)H2(Wr2p,Kr2p)H2(Wr2p)H_{2}(Z,P\cup W)\cong H_{2}(-W_{r^{2}p},K_{r^{2}p})\cong H^{2}(-W_{r^{2}p})\cong\mathbb{Z}

So, the exact sequence simlpifies to

0H2(PW)H2(Z)0\rightarrow H_{2}(P\cup W)\rightarrow H_{2}(Z)\rightarrow\mathbb{Z}\rightarrow\dots

Since kzkz is in H2(PW)H_{2}(P\cup W), it gets mapped to 0 under the map H2(Z)H_{2}(Z)\rightarrow\mathbb{Z}. Since \mathbb{Z} is free, zz also gets mapped to 0 under the map H2(Z)H_{2}(Z)\rightarrow\mathbb{Z}. Hence, zH2(PW)z\in H_{2}(P\cup W).

Since H2(P)H2(W)H2(PW)H_{2}(P)\oplus H_{2}(W)\rightarrow H_{2}(P\cup W) is an isomorphism, we must have z=p+τz=p+\tau for some pH2(P)p\in H_{2}(P) and τH2(W)\tau\in H_{2}(W). ∎

Lemma 3.3.

In the intersection form of ZZ, H2(P)H_{2}(P) is the orthogonal copmlement of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}).

Proof.

From the Mayer-Vietoris sequence

0H2(P)H2(WWr2p)H2(Z)H1(L(p,q))0\rightarrow H_{2}(P)\oplus H_{2}(W\cup-W_{r^{2}p})\rightarrow H_{2}(Z)\rightarrow H_{1}(L(p,q))\rightarrow\dots

We know that H2(P)H2(WWr2p)H2(Z)H_{2}(P)\oplus H_{2}(W\cup-W_{r^{2}p})\rightarrow H_{2}(Z) is a full rank embedding. Since H2(P)H_{2}(P) and H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}) are orthogonal to each other, Lemma 3.2 implies that H2(P)H_{2}(P) is the orthogonal complement of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}). ∎

Lemma 3.4.

The image of the map H2(W)H2(WWr2p)H_{2}(W)\rightarrow H_{2}(W\cup-W_{r^{2}p}) induced by inclusion is exactly the torsion subgroup of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}).

Proof.

By excision and the neighborhood collar theorem, we have H2(WWr2p,W)H2(Wr2p,Kr2p)H2(Wr2p)H_{2}(W\cup-W_{r^{2}p},W)\cong H_{2}(-W_{r^{2}p},K_{r^{2}p})\cong H^{2}(-W_{r^{2}p}), which has no torsion because H1(Wr2p)=0H_{1}(-W_{r^{2}p})=0.

Consider the exact sequence of a pair

H2(W)H2(WWr2p)H2(WWr2p,W)\rightarrow H_{2}(W)\rightarrow H_{2}(W\cup-W_{r^{2}p})\rightarrow H_{2}(W\cup-W_{r^{2}p},W)\rightarrow

Since H2(WWr2p,W)H_{2}(W\cup-W_{r^{2}p},W) is free, the image of the map H2(W)H2(WWr2p)H_{2}(W)\rightarrow H_{2}(W\cup-W_{r^{2}p}) must contains all torsion elements. Since H2(W)H_{2}(W) is finite, the image is exactly the torsion subgroup. ∎

Lemma 3.5.

The map H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) induced by the inclusion L(p,q)Kr2pWL(p,q)\cup K_{r^{2}p}\rightarrow W has kernal of size rprp and image of size rprp.

Proof.

Consider the exact sequence of a pair

0H2(W)H2(W,L(p,q)Kr2p)H1(L(p,q)Kr2p)H1(W)\displaystyle 0\rightarrow H_{2}(W)\rightarrow H_{2}(W,L(p,q)\cup K_{r^{2}p})\rightarrow H_{1}(L(p,q)\cup K_{r^{2}p})\rightarrow H_{1}(W)
H1(W,L(p,q)Kr2p)2\displaystyle\rightarrow H_{1}(W,L(p,q)\cup K_{r^{2}p})\rightarrow\mathbb{Z}^{2}\rightarrow\dots

Using Poincaré-Lefschetz duality and universal coefficient theorem, we get

0H2(W)H1(W)H1(L(p,q))H1(Kr2p)H1(W)\displaystyle 0\rightarrow H_{2}(W)\rightarrow H_{1}(W)\rightarrow H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W)
H2(W)2\displaystyle\rightarrow H_{2}(W)\oplus\mathbb{Z}\rightarrow\mathbb{Z}^{2}\rightarrow\dots

which simplifies to

0H2(W)H1(W)H1(L(p,q))H1(Kr2p)H1(W)H2(W)00\rightarrow H_{2}(W)\rightarrow H_{1}(W)\rightarrow H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W)\rightarrow H_{2}(W)\rightarrow 0

Hence, the map H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) has kernal of size |H1(W)||H2(W)|\dfrac{|H_{1}(W)|}{|H_{2}(W)|} and image of size |H1(W)||H2(W)|\dfrac{|H_{1}(W)|}{|H_{2}(W)|}. Since the kernal and the image has the same size, their sizes must both be |H1(L(p,q))H1(Kr2p)|=rp\sqrt{|H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})|}=rp. ∎

Lemma 3.6.

The kernel of the map H1(Kr2p)H1(W)H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) induced by inclusion is isomorphic to r\mathbb{Z}_{r} (cyclic group of order rr).

Proof.

Due to the surjectivitiy of the map labeled as “ss” in Lemma 2.3, we know that the image of the map H1(L(p,q))H1(W)H_{1}(L(p,q))\rightarrow H_{1}(W) is contained in the image of the map H1(Kr2p)H1(W)H_{1}(K_{r^{2}p})\rightarrow H_{1}(W). Hence, the image of the map H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) is the same as the image of the map H1(Kr2p)H1(W)H_{1}(K_{r^{2}p})\rightarrow H_{1}(W).

Therefore, by Lemma 3.5, the map H1(Kr2p)H1(W)H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) has an image of size rprp and a kernal of size r2prp=r\dfrac{r^{2}p}{rp}=r. Since H1(Kr2p)H_{1}(K_{r^{2}p}) is cyclic, the kernal is a cyclic group of order rr. ∎

Recall that H2(Wr2p)H_{2}(-W_{r^{2}p}) is generated by [Σ][\Sigma]. The follow Lemma is about the generator of the free part of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}):

Lemma 3.7.

The map H2(Wr2p)H2(WWr2p)H_{2}(-W_{r^{2}p})\rightarrow H_{2}(W\cup-W_{r^{2}p}) induced by inclusion maps [Σ][\Sigma] to rr times a generator of the free part of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}) plus a torsion element (can be 0).

Proof.

Consider the Mayer-Vietoris sequence

0H2(W)H2(Wr2p)H2(WWr2p)H1(Kr2p)H1(W)0\rightarrow H_{2}(W)\oplus H_{2}(-W_{r^{2}p})\rightarrow H_{2}(W\cup-W_{r^{2}p})\rightarrow H_{1}(K_{r^{2}p})\rightarrow H_{1}(W)\rightarrow\dots

By Lemma 3.6, we can simplify this into

0H2(W)H2(Wr2p)H2(WWr2p)r00\rightarrow H_{2}(W)\oplus H_{2}(-W_{r^{2}p})\rightarrow H_{2}(W\cup-W_{r^{2}p})\rightarrow\mathbb{Z}_{r}\rightarrow 0

Combining this with Lemma 3.4, we conclude that H2(Wr2p)H2(WWr2p)H_{2}(-W_{r^{2}p})\rightarrow H_{2}(W\cup-W_{r^{2}p}) has to map [Σ][\Sigma] to rr times a generator of the free part of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}) plus a torsion element (can be 0). ∎

Let σH2(WWr2p)\sigma\in H_{2}(W\cup-W_{r^{2}p}) be a generator of the free part such that [Σ]=rσ+torsion[\Sigma]=r\sigma+torsion.

Recall that the intersection pairing of [Σ][\Sigma] with itself gives r2p-r^{2}p. So we have

σ,σ=1r2[Σ],[Σ]=r2pr2=p\langle\sigma,\sigma\rangle=\dfrac{1}{r^{2}}\langle[\Sigma],[\Sigma]\rangle=\dfrac{-r^{2}p}{r^{2}}=-p

Since PP is negative definite, by Lemma 3.3, we know that ZZ is negative definite. By Donaldson’s Theorem [8], we know that the intersection pairing on ZZ is n+1-\mathbb{Z}^{n+1}, where nn is the rank of H2(P)H_{2}(P). Moreover, under the identification H2(Z)n+1H_{2}(Z)\cong-\mathbb{Z}^{n+1}, the image of the first Chern class map c1:Spinc(Z)H2(Z)H2(Z)n+1c_{1}:\text{Spin}^{\text{c}}(Z)\rightarrow H^{2}(Z)\cong H_{2}(Z)\cong-\mathbb{Z}^{n+1} is exactly the set of characteristic vectors

Char(n+1):={vn+1|wn+1v,ww,w (mod 2)}\text{Char}(-\mathbb{Z}^{n+1}):=\{v\in-\mathbb{Z}^{n+1}|\forall w\in-\mathbb{Z}^{n+1}\ \langle v,w\rangle\equiv\langle w,w\rangle\text{ (mod 2)}\}

where H2(Z)H2(Z)H^{2}(Z)\cong H_{2}(Z) is Poincaré duality.

In any orthonormal basis, the set of characteristic vectors of n+1-\mathbb{Z}^{n+1} is exactly the set of vectors with all entries being odd.

Now we can prove the following:

Lemma 3.8.

A spinc\text{spin}^{\text{c}} structure on Kr2pK_{r^{2}p} extends to WW if and only if the label ii on the spinc\text{spin}^{\text{c}} structure satisfies i{r2(modr) if r is even and p is odd0(modr) otherwisei\equiv\begin{cases}\dfrac{r}{2}\quad(mod\ r)\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\quad(mod\ r)\text{ otherwise}\end{cases}.

Note that since PP and Wr2p-W_{r^{2}p} have trivial H1H_{1}, a spinc\text{spin}^{\text{c}} structure on on Kr2pK_{r^{2}p} extends to WW if and only if it extends to ZZ. So, in the statement of Lemma 3.8, we can always replace the WW with ZZ.

Proof of Lemma 3.8.

Consider the map H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) mentioned in Lemma 3.5. By Lemma 3.5, it has a kernal of size rprp. Since the map H1(L(p,q))H1(W)H_{1}(L(p,q))\rightarrow H_{1}(W) is injective, the kernal of H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) cannot contain two different elements with the same H1(Kr2p)H_{1}(K_{r^{2}p}) component. Hence, the rprp elements in the kernal of H1(L(p,q))H1(Kr2p)H1(W)H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W) all have different H1(Kr2p)H_{1}(K_{r^{2}p}) components.

Therefore, by applying Poincaré-Lefschetz duality on the exact sequence of a pair

H2(W,L(p,q)Kr2p)H1(L(p,q))H1(Kr2p)H1(W)\dots\rightarrow H_{2}(W,L(p,q)\cup K_{r^{2}p})\rightarrow H_{1}(L(p,q))\oplus H_{1}(K_{r^{2}p})\rightarrow H_{1}(W)\rightarrow\dots

we can conclude that the map H2(W)H2(Kr2p)H^{2}(W)\rightarrow H^{2}(K_{r^{2}p}) has image of size rprp. Therefore there are exactly rprp spinc\text{spin}^{\text{c}} structures on Kr2pK_{r^{2}p} that extend to WW.

Notice that there are exactly rprp possible values of i(modr2p)i\ (mod\ r^{2}p) satisfying

(1) i{r2(modr) if r is even and p is odd0(modr) otherwisei\equiv\begin{cases}\dfrac{r}{2}\quad(mod\ r)\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\quad(mod\ r)\text{ otherwise}\end{cases}

Therefore, to prove Lemma 3.8, it suffices to show that a spinc\text{spin}^{\text{c}} structure on Kr2pK_{r^{2}p} extends to WW implies that its label ii satisfies (1). By the definition of the labeling, this is equivalent to showing that for any spinc\text{spin}^{\text{c}} structure 𝔰\mathfrak{s} on ZZ,

c1(𝔰),[Σ]+r2p{r(mod 2r) if r is even and p is odd0(mod 2r) otherwise\langle c_{1}(\mathfrak{s}),[\Sigma]\rangle+r^{2}p\equiv\begin{cases}r\quad(mod\ 2r)\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\quad(mod\ 2r)\text{ otherwise}\end{cases}

Since c1(𝔰),[Σ]+r2p=r(c1(𝔰),σ+rp)\langle c_{1}(\mathfrak{s}),[\Sigma]\rangle+r^{2}p=r(\langle c_{1}(\mathfrak{s}),\sigma\rangle+rp), we only need to figure out whether c1(𝔰),σ+rp\langle c_{1}(\mathfrak{s}),\sigma\rangle+rp is odd or even.

Since c1(𝔰)c_{1}(\mathfrak{s}) is a characteristic vector, we have

c1(𝔰),σσ,σ=p(mod 2)\langle c_{1}(\mathfrak{s}),\sigma\rangle\equiv\langle\sigma,\sigma\rangle=-p\quad(mod\ 2)

Hence, c1(𝔰),σ+rp\langle c_{1}(\mathfrak{s}),\sigma\rangle+rp is odd when rr is even and pp is odd, and it is even otherwise. ∎

By [9, Th. 1.7] and Lemma 3.3, to prove Theorem 1.5, we only need to show that σ\sigma embeds as a changemaker vector in the lattice n+1-\mathbb{Z}^{n+1}.

By choosing suitable orthonormal basis, we let σ=(σ0,,σn)\sigma=(\sigma_{0},\dots,\sigma_{n}) with

0σnσ00\leq\sigma_{n}\leq\dots\leq\sigma_{0}

For every knot KK in S3S^{3}, based on work by Rasmussen [15] and also Ni and Wu [14], there is a sequence of non-negative integers Vi(K)V_{i}(K) satisfying the following properties:

  • i0\forall i\geq 0, Vi(K)Vi+1(K){0,1}V_{i}(K)-V_{i+1}(K)\in\{0,1\}

  • Vg(K)(K)=0V_{g(K)}(K)=0, where g(K)g(K) is the genus of the knot

  • For all positive integer kk, for all ii with 0ik10\leq i\leq k-1, we have

    2Vmin{i,ki}=d(Kk,i)d(Uk,i)-2V_{\text{min}\{i,k-i\}}=d(K_{k},i)-d(U_{k},i)

    where KkK_{k} is the kk-surgery on KK, and UkU_{k} is the kk-surgery on the unknot (which is L(k,k1)L(k,k-1)). The ii behind denotes the spinc\text{spin}^{\text{c}} structure with label ii, and the dd is the dd-invariant.

The last property implies that for 0ir2p20\leq i\leq\dfrac{r^{2}p}{2}, we have

2Vi(K)=d(Kr2p,i)d(Ur2p,i)-2V_{i}(K)=d(K_{r^{2}p},i)-d(U_{r^{2}p},i)

The smooth concordance invariant ν+(K)\nu^{+}(K) is the smallest ii such that Vi(K)=0V_{i}(K)=0. It is a lower bound of the 4-ball genus of KK [11, Prop. 2.4]. When KK is an L-space knot, ν+\nu^{+} coincides with the genus of KK and these VV coefficients coincide with the torsion coefficients that can be computed from the Alexander polynomial.

For notational simplicity, we often drop the (K)(K) and only write ViV_{i} and ν+\nu^{+}.

We now proceed to prove that σ\sigma is a changemaker vector. Our proof will follow the argument in [10, §2-3] that shows that σ\sigma is a changemaker. The only difference is that we use the VV coefficients instead of torsion coefficients because we are not assuming KK to be an L-space knot, and there are some other minor changes because of the rr factor coming from our cobordism which did not exist in [10, §2-3].

Based on Lemma 3.8, we make the following definition:

Definition 3.9.

We say that the coefficient ViV_{i} is a relevant coefficient if the index satisfies 0ir2p20\leq i\leq\dfrac{r^{2}p}{2} and

i{r2(modr) if r is even and p is odd0(modr) otherwisei\equiv\begin{cases}\dfrac{r}{2}\quad(mod\ r)\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\quad(mod\ r)\text{ otherwise}\end{cases}

In that case, we say that the index ii is relevant.

Lemma 3.10.

Let ii be relevant. Then

8Vi=max𝔠Char(n+1)𝔠,[Σ]+r2p=2i(𝔠2+(n+1))-8V_{i}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(-\mathbb{Z}^{n+1})\\ \langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i\end{subarray}}(\mathfrak{c}^{2}+(n+1))
Proof of Lemma 3.10.

Let ii be relevant. There is a spinc\text{spin}^{\text{c}} structure on Kr2pK_{r^{2}p} with label ii. By Lemma 3.8, this spinc\text{spin}^{\text{c}} structure extends to some spinc\text{spin}^{\text{c}} structure 𝔰iW\mathfrak{s}^{W}_{i} on WW. Let 𝔰iL(p,q)\mathfrak{s}^{L(p,q)}_{i} be the restriction of 𝔰iW\mathfrak{s}^{W}_{i} on L(p,q)L(p,q). (Warning: In the literature, there is a convention of labeling spinc\text{spin}^{\text{c}} structures on lens spaces with numbers 0,,p10,\dots,p-1. The label of 𝔰iL(p,q)\mathfrak{s}^{L(p,q)}_{i} under such a convention may not necessarily be ii.)

Now we recreate [10, Lemma 2.3], but under the context of this paper. Since the canonical negative plumbing of a lens space is always sharp (see [10, Lemma 2.1]), the argument in the proof of [10, Lemma 2.3] implies that for any spinc\text{spin}^{\text{c}} structure 𝔰iP\mathfrak{s}^{P}_{i} on PP that extends 𝔰iL(p,q)\mathfrak{s}^{L(p,q)}_{i}, we have

c1(𝔰iP)2+n4d(𝔰iL(p,q))c_{1}(\mathfrak{s}^{P}_{i})^{2}+n\leq 4d(\mathfrak{s}^{L(p,q)}_{i})

And for any spinc\text{spin}^{\text{c}} structure 𝔰iWr2p\mathfrak{s}^{-W_{r^{2}p}}_{i} on Wr2p-W_{r^{2}p} that extends (Kr2p,i)(K_{r^{2}p},i), we have

c1(𝔰iWr2p)2+14d(Ur2p,i)c_{1}(\mathfrak{s}^{-W_{r^{2}p}}_{i})^{2}+1\leq-4d(U_{r^{2}p},i)

Furthermore, 𝔰iP\mathfrak{s}^{P}_{i} and 𝔰iWr2p\mathfrak{s}^{-W_{r^{2}p}}_{i} can be chosen such that equality holds. Adding these two inequalities together, we have

c1(𝔰iP)2+c1(𝔰iWr2p)2+(n+1)4d(𝔰iL(p,q))4d(Ur2p,i)c_{1}(\mathfrak{s}^{P}_{i})^{2}+c_{1}(\mathfrak{s}^{-W_{r^{2}p}}_{i})^{2}+(n+1)\leq 4d(\mathfrak{s}^{L(p,q)}_{i})-4d(U_{r^{2}p},i)

Now we look at the cobordism WW that exists in this paper but not in [10]. Since the dd-invariant is invariant under smooth rational homology cobordism, we know that

d(𝔰iL(p,q))=d(Kr2p,i)d(\mathfrak{s}^{L(p,q)}_{i})=d(K_{r^{2}p},i)

Also, since L(p,q)L(p,q) and Kr2pK_{r^{2}p} are rational homology spheres and H2(W)H_{2}(W) is finite, we have

c1(𝔰)2=c1(𝔰iP)2+c1(𝔰iWr2p)2c_{1}(\mathfrak{s})^{2}=c_{1}(\mathfrak{s}^{P}_{i})^{2}+c_{1}(\mathfrak{s}^{-W_{r^{2}p}}_{i})^{2}

Hence, for any spinc\text{spin}^{\text{c}} structure 𝔰\mathfrak{s} on ZZ that extends (Kr2p,i)(K_{r^{2}p},i), we have

c1(𝔰)2+(n+1)4d(Kr2p,i)4d(Ur2p,i)=8Vic_{1}(\mathfrak{s})^{2}+(n+1)\leq 4d(K_{r^{2}p},i)-4d(U_{r^{2}p},i)=-8V_{i}

and for any relevant ii, 𝔰\mathfrak{s} can be chosen such that equality holds.

Due to the way the labeling of spinc\text{spin}^{\text{c}} structures on Kr2pK_{r^{2}p} works, we know that spinc\text{spin}^{\text{c}} structure 𝔰\mathfrak{s} on ZZ that extends (Kr2p,i)(K_{r^{2}p},i) are exactly the ones that satisfies

(2) c1(𝔰),[Σ]+r2p2i (mod 2r2p)\langle c_{1}(\mathfrak{s}),[\Sigma]\rangle+r^{2}p\equiv 2i\text{ (mod }2r^{2}p\text{)}

Hence, we conclude that

(3) 8Vi=max𝔠Char(n+1)𝔠,[Σ]+r2p2i (mod 2r2p)(𝔠2+(n+1))-8V_{i}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(-\mathbb{Z}^{n+1})\\ \langle\mathfrak{c},[\Sigma]\rangle+r^{2}p\equiv 2i\text{ (mod }2r^{2}p\text{)}\end{subarray}}(\mathfrak{c}^{2}+(n+1))

(3) is very close to the statement of Lemma 3.10. The only difference is that in Lemma 3.10, it says “=2i=2i” instead of “2i\equiv 2i (mod 2r2p2r^{2}p)” under the max\max.

To finish the proof of Lemma 3.10, it suffices to show that for any 𝔠=c1(𝔰)Char(n+1)\mathfrak{c}=c_{1}(\mathfrak{s})\in Char(-\mathbb{Z}^{n+1}) satisfying (2), there exists some 𝔠\mathfrak{c}^{\prime} with 𝔠2𝔠2{\mathfrak{c}^{\prime}}^{2}\geq\mathfrak{c}^{2} satisfying

𝔠,[Σ]+r2p=2i\langle\mathfrak{c}^{\prime},[\Sigma]\rangle+r^{2}p=2i

To do this, we apply the argument in the proof of [12, Lemma 2.9]. Let 𝔠=c1(𝔰)Char(n+1)\mathfrak{c}=c_{1}(\mathfrak{s})\in Char(-\mathbb{Z}^{n+1}) satisfying (2). Then

𝔠,[Σ]+r2p=2i+2mr2p\langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i+2mr^{2}p

for some mm\in\mathbb{Z}.

Let 𝔠:=𝔠+2m[Σ]\mathfrak{c}^{\prime}:=\mathfrak{c}+2m[\Sigma]. Then we have

𝔠,[Σ]+r2p=2i+2mr2p+2m[Σ],[Σ]=2i\langle\mathfrak{c}^{\prime},[\Sigma]\rangle+r^{2}p=2i+2mr^{2}p+2m\langle[\Sigma],[\Sigma]\rangle=2i

And also

𝔠2\displaystyle{\mathfrak{c}^{\prime}}^{2} =𝔠2+4m𝔠,[Σ]+4m2[Σ],[Σ]\displaystyle=\mathfrak{c}^{2}+4m\langle\mathfrak{c},[\Sigma]\rangle+4m^{2}\langle[\Sigma],[\Sigma]\rangle
=𝔠2+4m(2i+2mr2pr2p)+4m2(r2p)\displaystyle=\mathfrak{c}^{2}+4m(2i+2mr^{2}p-r^{2}p)+4m^{2}(-r^{2}p)
=𝔠2+4r2pm2+(8i4r2p)m\displaystyle=\mathfrak{c}^{2}+4r^{2}pm^{2}+(8i-4r^{2}p)m

Viewing 4r2pm2+(8i4r2p)m4r^{2}pm^{2}+(8i-4r^{2}p)m as a quadratic in mm, the roots are 0 and (12ir2p)(1-\dfrac{2i}{r^{2}p}). Since ii is relevant, we have 02ir2p0\leq 2i\leq r^{2}p. Hence, there is no integers strictly between 0 and (12ir2p)(1-\dfrac{2i}{r^{2}p}). Therefore, since mm is an integer, we have 4r2pm2+(8i4r2p)m04r^{2}pm^{2}+(8i-4r^{2}p)m\geq 0, and hence 𝔠2𝔠2{\mathfrak{c}^{\prime}}^{2}\geq\mathfrak{c}^{2}. ∎

Lemma 3.11.

σ\sigma is a changemaker vector.

Proof.

First of all, similar to the argument in [9, §3.4], by Lemma 3.3 and the indecomposability of linear lattices [9, Cor. 3.5], we know that all entries of σ\sigma are non-zero.

When 𝔠=(1,,1)\mathfrak{c}=(1,\dots,1), we have

𝔠,[Σ]+r2p\displaystyle\langle\mathfrak{c},[\Sigma]\rangle+r^{2}p =r𝔠,σ+r2p\displaystyle=r\langle\mathfrak{c},\sigma\rangle+r^{2}p
=r|σ|1+r2p\displaystyle=-r|\sigma|_{1}+r^{2}p
=2(r2pr|σ|12)\displaystyle=2\left(\dfrac{r^{2}p-r|\sigma|_{1}}{2}\right)

Also, 𝔠2+(n+1)=0\mathfrak{c}^{2}+(n+1)=0.

By Lemma 3.10, since those ViV_{i} coefficients are all non-negative, we can conclude that

V12r(rp|σ|1)=0V_{\tfrac{1}{2}r(rp-|\sigma|_{1})}=0

Therefore, ν+12r(rp|σ|1)\nu^{+}\leq\tfrac{1}{2}r(rp-|\sigma|_{1}). Hence, for all i12r(rp|σ|1)i\geq\tfrac{1}{2}r(rp-|\sigma|_{1}), we have Vi=0V_{i}=0.

By Lemma 3.10, we know that for all relevant i12r(rp|σ|1)i\geq\tfrac{1}{2}r(rp-|\sigma|_{1}), there exists some 𝔠Char(n+1)\mathfrak{c}\in Char(-\mathbb{Z}^{n+1}) such that 𝔠2=(n+1)\mathfrak{c}^{2}=-(n+1) and 𝔠,[Σ]+r2p=2i\langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i. This is equivalent to saying that there exists some 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that

r𝔠,σ+r2p=2ir\langle\mathfrak{c},\sigma\rangle+r^{2}p=2i

For the rest of the proof, we will separate the cases according to Lemma 3.8. The proofs for the two cases are almost identical.

Case 1: rr is odd or pp is even
Let ii be relevant and i12r(rp|σ|1)i\geq\tfrac{1}{2}r(rp-|\sigma|_{1}). When rr is odd or pp is even, that is equivalent to saying that ii is a multiple of rr that satisfies

12r(rp|σ|1)ir2p2\tfrac{1}{2}r(rp-|\sigma|_{1})\leq i\leq\tfrac{r^{2}p}{2}

Let i:=iri^{\prime}:=\dfrac{i}{r}.

Then we have rp|σ|12irp2\tfrac{rp-|\sigma^{\prime}|_{1}}{2}\leq i^{\prime}\leq\tfrac{rp}{2}, and there exists 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that

r𝔠,σ+r2p=2irr\langle\mathfrak{c},\sigma\rangle+r^{2}p=2i^{\prime}r

which simplifies to

(4) 𝔠,σ+rp=2i\langle\mathfrak{c},\sigma\rangle+rp=2i^{\prime}

Let i′′:=rpii^{\prime\prime}:=rp-i^{\prime}. Then we have rp2i′′rp+|σ|12\tfrac{rp}{2}\leq i^{\prime\prime}\leq\tfrac{rp+|\sigma|_{1}}{2}. Substituting i=rpi′′i^{\prime}=rp-i^{\prime\prime} into (4) and rearranging, we get

𝔠,σ+rp=2i′′\langle-\mathfrak{c},\sigma\rangle+rp=2i^{\prime\prime}

Notice that 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} implies 𝔠{±1}n+1-\mathfrak{c}\in\{\pm 1\}^{n+1}. Hence, we can conclude that for any integer ii^{\prime} satisfying rp|σ|12irp+|σ|12\tfrac{rp-|\sigma^{\prime}|_{1}}{2}\leq i^{\prime}\leq\tfrac{rp+|\sigma^{\prime}|_{1}}{2}, there exists 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that (4) holds.

Note that since |σ|1σ,σ=p|\sigma|_{1}\equiv\langle\sigma,\sigma\rangle=-p (mod 22), rp±|σ|12\tfrac{rp\pm|\sigma|_{1}}{2} must be integers under Case 1.

Setting j:=irp+|σ|12j:=i^{\prime}-\tfrac{rp+|\sigma|_{1}}{2} and substituting it into (4), we arrive to the conclusion that for any integer jj satisfying |σ|1j0-|\sigma|_{1}\leq j\leq 0, there exists 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that

𝔠,σ=2j+|σ|1\langle\mathfrak{c},\sigma\rangle=2j+|\sigma|_{1}

Setting 𝔠=(1,,1)+2χ\mathfrak{c}=(-1,\dots,-1)+2\chi for some χ{0,1}n+1\chi\in\{0,1\}^{n+1}, we arrive to the conclusion that for any integer jj satisfying |σ|1j0-|\sigma|_{1}\leq j\leq 0, there exists χ{0,1}n+1\chi\in\{0,1\}^{n+1} such that

χ,σ=j\langle\chi,\sigma\rangle=j

Together with the fact that all entries of σ\sigma are non-zero, this implies σ\sigma being a changemaker vector.

Case 2: rr is even and pp is odd
Let ii be relevant and i12r(rp|σ|1)i\geq\tfrac{1}{2}r(rp-|\sigma|_{1}). When rr is even and pp is odd, that is equivalent to saying that ii satisfies ir2i\equiv\tfrac{r}{2} (mod rr) and

12r(rp|σ|1)ir2p2\tfrac{1}{2}r(rp-|\sigma|_{1})\leq i\leq\tfrac{r^{2}p}{2}

Let i:=ir12i^{\prime}:=\tfrac{i}{r}-\tfrac{1}{2}.

Then we have rp|σ|12i+12rp2\tfrac{rp-|\sigma^{\prime}|_{1}}{2}\leq i^{\prime}+\tfrac{1}{2}\leq\tfrac{rp}{2}, and there exists 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that

r𝔠,σ+r2p=2(i+12)rr\langle\mathfrak{c},\sigma\rangle+r^{2}p=2\left(i^{\prime}+\tfrac{1}{2}\right)r

which simplifies to

(5) 𝔠,σ+rp=2i+1\langle\mathfrak{c},\sigma\rangle+rp=2i^{\prime}+1

Let i′′:=rpi1i^{\prime\prime}:=rp-i^{\prime}-1. Then rp2i′′+12rp+|σ|12\tfrac{rp}{2}\leq i^{\prime\prime}+\tfrac{1}{2}\leq\tfrac{rp+|\sigma|_{1}}{2} and

𝔠,σ+rp=2i′′+1\langle-\mathfrak{c},\sigma\rangle+rp=2i^{\prime\prime}+1

Hence (5) holds for rp|σ|12i′′+12rp+|σ|12\tfrac{rp-|\sigma|_{1}}{2}\leq i^{\prime\prime}+\tfrac{1}{2}\leq\tfrac{rp+|\sigma|_{1}}{2}.

Note that in Case 2, rp+|σ|1rp+|\sigma|_{1} is odd. Therefore, rp+|σ|112\tfrac{rp+|\sigma|_{1}-1}{2} is an integer.

Setting j:=irp+|σ|112j:=i^{\prime}-\tfrac{rp+|\sigma^{\prime}|_{1}-1}{2} and substituting it into (5), we see that for any integer jj satisfying |σ|1j0-|\sigma^{\prime}|_{1}\leq j\leq 0, there exists 𝔠{±1}n+1\mathfrak{c}\in\{\pm 1\}^{n+1} such that

𝔠,σ=2j+|σ|1\langle\mathfrak{c},\sigma\rangle=2j+|\sigma|_{1}

The rest of the proof is the same as Case 1. ∎

Proof of Theorem 1.5.

Theorem 1.5 follows from [9, Th. 1.7], Lemma 3.3, and Lemma 3.11. ∎

4. Number of possible surgeries when fixing rr

In this section we will show that when we fix KS3K\subset S^{3} and a value of rr, the number of (r,p)(r,p)-lensbordant surgeries is very limited. We will follow an argument similar to [12, §2], with some changes to take rr into account.

Recall that we defined the coefficient ViV_{i} to relevant if 0ir2p20\leq i\leq\dfrac{r^{2}p}{2} and

i{r2(modr) if r is even and p is odd0(modr) otherwisei\equiv\begin{cases}\dfrac{r}{2}\quad(mod\ r)\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\quad(mod\ r)\text{ otherwise}\end{cases}

Sometimes it is convenient to consider the relevant coefficients as a sequence itself:

Definition 4.1.

For ii satisfying

0i{rp21 if r is even and p is oddrp12 if r,p are both oddrp2 if p is even0\leq i\leq\begin{cases}\frac{rp}{2}-1\text{ if }r\text{ is even and }p\text{ is odd}\\ \frac{rp-1}{2}\text{ if }r,p\text{ are both odd}\\ \frac{rp}{2}\text{ if }p\text{ is even}\end{cases}

we define

Virel:={Vr2+ir if r is even and p is oddVir otherwiseV^{rel}_{i}:=\begin{cases}V_{\frac{r}{2}+ir}\text{ if }r\text{ is even and }p\text{ is odd}\\ V_{ir}\text{ otherwise}\end{cases}

This range of ii for VirelV^{rel}_{i} corresponds to the range of ii for ViV_{i} to be relevant.

Since the VV coefficients of a knot are always monotonic decreasing, the relevant coefficients are also monotonic decreasing. However, consecutive relevant coefficients can differ by more than 1.

We define ν+rel\nu^{+rel} to be the number of non-zero relevant coefficients. Equivalently,

ν+rel:=min{i|Virel=0}\nu^{+rel}:=min\{i|V^{rel}_{i}=0\}

Similar to how McCoy [12] defined TmT_{m}, we define TmrelT^{rel}_{m} be the number of relevant coefficients less than or equal to mm. Equivalently,

Tmrel:=|{0i<ν+rel0<Virelm}|T^{rel}_{m}:=|\{0\leq i<\nu^{+rel}\mid 0<V^{rel}_{i}\leq m\}|

Note that due to monotonicity, we have ν+rel=TV0relrel\nu^{+rel}=T^{rel}_{V^{rel}_{0}}.

We define Sm={α0dim(σ)jαj(αj+1)=2m}S_{m}=\{\alpha\in\mathbb{Z}^{dim(\sigma)}_{\geq 0}\mid\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)=2m\}. This is exactly the same as how McCoy defined SmS_{m} in [12].

We also define Sm={α0dim(σ)jαj(αj+1)2m}S_{\leq m}=\{\alpha\in\mathbb{Z}^{dim(\sigma)}_{\geq 0}\mid\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\leq 2m\}. Due to the fact that consecutive relevant coefficients can decrease by more than 1, some arguments later in this section will not work if we use SmS_{m} like how it is used in [12]. Instead, we will need to consider SmS_{\leq m}.

First we prove the following properties, which is an analog of [12, Lemma 2.10]:

Lemma 4.2.

The following three statements are true:
(1) For all mm satisfying 0m<V0rel0\leq m<V^{rel}_{0}, we have Tmrel=maxαSmσαT^{rel}_{m}=\displaystyle\max_{\alpha\in S_{\leq m}}\sigma\cdot\alpha

(2) ν+rel={12+12(rp|σ|1) if r is even and p is odd12(rp|σ|1) otherwise\nu^{+rel}=\begin{cases}-\dfrac{1}{2}+\dfrac{1}{2}(rp-|\sigma|_{1})\text{ if }r\text{ is even and }p\text{ is odd}\\ \dfrac{1}{2}(rp-|\sigma|_{1})\text{ otherwise}\end{cases}

(3) ν+rel=TV0relrelmaxαSV0relσα\nu^{+rel}=T^{rel}_{V^{rel}_{0}}\leq\displaystyle\max_{\alpha\in S_{V^{rel}_{0}}}\sigma\cdot\alpha

Note that when r2r\geq 2, statement (1) holds for m=V0relm=V^{rel}_{0} as well. We will prove this in Section 6 (Corollary 6.13) after we study more about the behavior of relevant coefficients.

Proof of Lemma 4.2.

Lemma 3.10 states when for any relevant ii,

8Vi=max𝔠Char(n+1)𝔠,[Σ]+r2p=2i(𝔠2+(n+1))-8V_{i}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(-\mathbb{Z}^{n+1})\\ \langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i\end{subarray}}(\mathfrak{c}^{2}+(n+1))

which can be rewritten as

8Virel\displaystyle-8V^{rel}_{i} =max𝔠Char(n+1)r𝔠,σ+r2p=2ir+r(𝔠2+(n+1)) if r is even and p is odd\displaystyle=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(-\mathbb{Z}^{n+1})\\ r\langle\mathfrak{c},\sigma\rangle+r^{2}p=2ir+r\end{subarray}}(\mathfrak{c}^{2}+(n+1))\text{ if }r\text{ is even and }p\text{ is odd}
8Virel\displaystyle-8V^{rel}_{i} =max𝔠Char(n+1)r𝔠,σ+r2p=2ir(𝔠2+(n+1)) otherwise\displaystyle=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(-\mathbb{Z}^{n+1})\\ r\langle\mathfrak{c},\sigma\rangle+r^{2}p=2ir\end{subarray}}(\mathfrak{c}^{2}+(n+1))\text{ otherwise}

The condition of 𝔠\mathfrak{c} being a characteristic vector can be stated as 𝔠=2α+(1,,1)\mathfrak{c}=2\alpha+(1,\dots,1) for some αn+1\alpha\in\mathbb{Z}^{n+1}. Now we can rewrite the above as

8Virel\displaystyle-8V^{rel}_{i} =maxrp|σ|12jσjαj=2i+1((n+1)j((2αj+1)2)) if r is even and p is odd\displaystyle=\max_{rp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=2i+1}((n+1)-\sum\limits_{j}((2\alpha_{j}+1)^{2}))\text{ if }r\text{ is even and }p\text{ is odd}
8Virel\displaystyle-8V^{rel}_{i} =maxrp|σ|12jσjαj=2i((n+1)j((2αj+1)2)) otherwise\displaystyle=\max_{rp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=2i}((n+1)-\sum\limits_{j}((2\alpha_{j}+1)^{2}))\text{ otherwise}

This can further simplify to

(6) 2Virel\displaystyle 2V^{rel}_{i} =minrp|σ|12jσjαj=2i+1jαj(αj+1) if r is even and p is odd\displaystyle=\min_{rp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=2i+1}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\text{ if }r\text{ is even and }p\text{ is odd}
2Virel\displaystyle 2V^{rel}_{i} =minrp|σ|12jσjαj=2ijαj(αj+1) otherwise\displaystyle=\min_{rp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=2i}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\text{ otherwise}

For any mm with 0m<V0rel0\leq m<V^{rel}_{0}, consider the minimal integer ii such that there exists α\alpha that satisfies

(7) rp|σ|12jσjαj={2i+1 if r is even and p is odd2i otherwiserp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=\begin{cases}2i+1\text{ if }r\text{ is even and }p\text{ is odd}\\ 2i\text{ otherwise}\end{cases}

and

jαj(αj+1)2m\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\leq 2m

Notice that all entries of such an α\alpha must be non-negative because if there is any negative entry αj\alpha_{j}, we can replace it with 1αj-1-\alpha_{j} which lowers ii but doesn’t change jαj(αj+1)\sum\limits_{j}\alpha_{j}(\alpha_{j}+1). Hence, such an α\alpha must be in SmS_{\leq m}. (Note that the entries being non-negative is in the definition of αSm\alpha\in S_{\leq m})

From the argument above, we see that if such an ii is non-negative, we have i=min{tVtrelm}i=min\{t\mid V^{rel}_{t}\leq m\}.

Now we argue that m<V0relm<V^{rel}_{0} implies such an ii being non-negative: Assume the contrary and suppose such a minimal ii is negative. Let kk be the smallest positive integer such that there exists αSm\alpha\in S_{\leq m} with

rp|σ|12jσjαj={2(k)+1 if r is even and p is odd2(k) otherwiserp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=\begin{cases}2(-k)+1\text{ if }r\text{ is even and }p\text{ is odd}\\ 2(-k)\text{ otherwise}\end{cases}

Let ll be maximal such that αl0\alpha_{l}\neq 0. If k>σlk>\sigma_{l}, we can decrease αl\alpha_{l} by 1 to produce a new α\alpha that is still in SmS_{\leq m} but kk decreases by σl\sigma_{l}, contradicting with the minimality of kk. Hence, we know that kσlk\leq\sigma_{l}. Since σ\sigma is a changemaker, A{l,,n}\exists A\subseteq\{l,\dots,n\} such that k=jAσjk=\sum\limits_{j\in A}\sigma_{j}. Consider α\alpha^{\prime} where

αj:={αj1 if jAαj otherwise\alpha^{\prime}_{j}:=\begin{cases}\alpha_{j}-1\text{ if }j\in A\\ \alpha_{j}\text{ otherwise}\end{cases}

Then α1n+1\alpha^{\prime}\in\mathbb{Z}_{\geq-1}^{n+1} satisfies jαj(αj+1)m\sum\limits_{j}\alpha^{\prime}_{j}(\alpha^{\prime}_{j}+1)\leq m and

rp|σ|12jσjαj={1 if r is even and p is odd0 otherwiserp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha^{\prime}_{j}=\begin{cases}1\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\text{ otherwise}\end{cases}

(6) implies that V0relmV_{0}^{rel}\leq m, which contradicts with m<V0relm<V^{rel}_{0}. Hence, we can conclude that m<V0relm<V^{rel}_{0} implies such an ii being non-negative, and therefore we have i=min{tVtrelm}i=min\{t\mid V^{rel}_{t}\leq m\}. So, we can conclude that

min{tVtrelm}={minαSm12(rp|σ|112jσjαj) if r is even and p is oddminαSm12(rp|σ|12jσjαj) otherwisemin\{t\mid V^{rel}_{t}\leq m\}=\begin{cases}\displaystyle\min_{\alpha\in S_{\leq m}}\frac{1}{2}\left(rp-|\sigma|_{1}-1-2\sum\limits_{j}\sigma_{j}\alpha_{j}\right)\text{ if }r\text{ is even and }p\text{ is odd}\\ \displaystyle\min_{\alpha\in S_{\leq m}}\frac{1}{2}\left(rp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}\right)\text{ otherwise}\end{cases}

Setting m=0m=0 gives statement (2) of Lemma 4.2. By monotonicity of relevant coefficients, we know that Tmrel=ν+relmin{tVtrelm}T^{rel}_{m}=\nu^{+rel}-min\{t\mid V^{rel}_{t}\leq m\}. That gives statement (1) of Lemma 4.2.

This argument does not fully apply to the m=V0relm=V^{rel}_{0} because of the issue of ii potentially being negative, causing it to not represent a relevant coefficient. However, we can still apply (6) to see that there exists some αn+1\alpha\in\mathbb{Z}^{n+1} such that 2V0rel=jαj(αj+1)2V^{rel}_{0}=\sum\limits_{j}\alpha_{j}(\alpha_{j}+1) and

rp|σ|12jσjαj={1 if r is even and p is odd0 otherwiserp-|\sigma|_{1}-2\sum\limits_{j}\sigma_{j}\alpha_{j}=\begin{cases}1\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\text{ otherwise}\end{cases}

We see that ν+rel=TV0relrel=σα\nu^{+rel}=T^{rel}_{V^{rel}_{0}}=\sigma\cdot\alpha for this particular α\alpha. Hence,

ν+rel=TV0relrelmaxαn+12V0rel=jαj(αj+1)σα\nu^{+rel}=T^{rel}_{V^{rel}_{0}}\leq\displaystyle\max_{\begin{subarray}{c}\alpha\in\mathbb{Z}^{n+1}\\ 2V^{rel}_{0}=\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\end{subarray}}\sigma\cdot\alpha

If α\alpha has any negative entry αj\alpha_{j}, we can replace it with 1αj-1-\alpha_{j} which increases σα\sigma\cdot\alpha but keeping jαj(αj+1)\sum\limits_{j}\alpha_{j}(\alpha_{j}+1) unchanged. Hence, we can take the max to be over αSV0rel\alpha\in S_{V^{rel}_{0}}, which gives statement (3) of Lemma 4.2. ∎

Remark 1: Since relevant coefficients can skip numbers, 2m2m may not equal to VirelV^{rel}_{i} for any ii, and hence we need to introduce SmS_{\leq m} in this paper instead of just SmS_{m} as in [12, §2].

Remark 2: The reason why the m=V0relm=V^{rel}_{0} case was dealt with separately in the proof is because of the issue of ii potentially being negative. If there is some condition that implies maxαSV0relσαmaxαSV0rel1σα\max\limits_{\alpha\in S_{V^{rel}_{0}}}\sigma\cdot\alpha-\max\limits_{\alpha\in S_{V^{rel}_{0}-1}}\sigma\cdot\alpha to be at most 1, since the minimal ii for m=V0rel1m=V^{rel}_{0}-1 in (7) is positive, such an ii for m=V0relm=V^{rel}_{0} must be non-negative. Hence, under such condition, statement (1) of Lemma 4.2 would also apply to m=V0relm=V^{rel}_{0}. In Section 6, we will show that r2r\geq 2 is such a condition.

Most of the remaining content of this section is for proving:

Theorem 1.7.

Let KK be a knot in S3S^{3} with ν+(K)0\nu^{+}(K)\neq 0. Then,
(1) When r=1r=1, there are at most 3 (r,p)(r,p)-lensbordant surgeries on KK.
(2) When rr is even, there are at most 2 (r,p)(r,p)-lensbordant surgeries on KK.
(3) When r3r\geq 3 is odd, there is at most 1 (r,p)(r,p)-lensbordant surgeries on KK.

Many ideas that go into the proof come from [12, §2].

First, we prove

Lemma 4.3.

Statement (1) of Theorem 1.7 is true.

Proof.

When r=1r=1, the conditions in Lemma 4.2 become the conditions in [12, Lemma 2.10]. Therefore, arguments in [12, §2-4] apply directly. ∎

For statement (2) and (3), notice that specifying KK and with the value of rr, and also specifying the parity of pp when rr is even, together determines all the relevant coefficients. So it suffices to show that the relevant coefficients and the value of rr, and the parity of pp when rr is even, together determines σ\sigma. That is because σ\sigma determines pp, which determines the surgery slope r2pr^{2}p when we know the value of rr.

We deal with the trivial cases first.

Lemma 4.4.

If V0rel=0V^{rel}_{0}=0 and r2r\geq 2, then r=2r=2 and σ=(1)\sigma=(1).

If V0rel=1V^{rel}_{0}=1 and r2r\geq 2, then it must be one of the following cases:
r=2r=2 and σ=(1,1)or(1,1,1)\sigma=(1,1)\ or\ (1,1,1)
r=3r=3 and σ=(1)\sigma=(1)
r=4r=4 and σ=(1)\sigma=(1)

If V0rel=2V^{rel}_{0}=2 and r2r\geq 2 and σ0=2\sigma_{0}=2 and σ1=1\sigma_{1}=1, then r=2r=2 and σ=(2,1)\sigma=(2,1).

Proof.

If V0rel=0V^{rel}_{0}=0, that means there are no non-zero relevant coefficients. Therefore

0=ν+rel={12+12(rp|σ|1) if r is even and p is odd12(rp|σ|1) otherwise0=\nu^{+rel}=\begin{cases}-\dfrac{1}{2}+\dfrac{1}{2}(rp-|\sigma|_{1})\text{ if }r\text{ is even and }p\text{ is odd}\\ \dfrac{1}{2}(rp-|\sigma|_{1})\text{ otherwise}\end{cases}

Since p|σ|1p\geq|\sigma|_{1}, given that r2r\geq 2, the only way this expression can be 0 is when r=2r=2 and p=|σ|1=1p=|\sigma|_{1}=1.

If V0rel=1V^{rel}_{0}=1, that means all non-zero relevant coefficients are 11. Hence we have ν+rel=T1rel\nu^{+rel}=T^{rel}_{1}. Now we have

0ν+rel=T1relmaxαS1σα=σ00\neq\nu^{+rel}=T^{rel}_{1}\leq\max_{\alpha\in S_{1}}\sigma\cdot\alpha=\sigma_{0}

Hence we have

{12+12(rσ02σ0)12+12(rp|σ|1)=grelσ0 if r is even and p is odd12(rσ02σ0)12(rp|σ|1)=grelσ0 otherwise\begin{cases}-\dfrac{1}{2}+\dfrac{1}{2}(r\sigma_{0}^{2}-\sigma_{0})\leq-\dfrac{1}{2}+\dfrac{1}{2}(rp-|\sigma|_{1})=g^{rel}\leq\sigma_{0}\text{ if }r\text{ is even and }p\text{ is odd}\\ \dfrac{1}{2}(r\sigma_{0}^{2}-\sigma_{0})\leq\dfrac{1}{2}(rp-|\sigma|_{1})=g^{rel}\leq\sigma_{0}\text{ otherwise}\end{cases}

Rearranging, we have

rσ0{3+1σ0 if r is even and p is odd3 otherwiser\sigma_{0}\leq\begin{cases}3+\frac{1}{\sigma_{0}}\text{ if }r\text{ is even and }p\text{ is odd}\\ 3\text{ otherwise}\end{cases}

So, we have r4r\leq 4. Given that r2r\geq 2, we also have σ0=1\sigma_{0}=1. So we can conclude that ν+rel1\nu^{+rel}\leq 1 and all entries of σ\sigma are 11. Since ν+rel0\nu^{+rel}\neq 0, we have ν+rel=1\nu^{+rel}=1. By considering the expression for ν+rel\nu^{+rel} in terms of r,p,|σ|1r,p,|\sigma|_{1}, we can see that the only possibilities are as stated in the lemma.

If V0rel=2V^{rel}_{0}=2 and r2r\geq 2 and σ0=2\sigma_{0}=2 and σ1=1\sigma_{1}=1, then

ν+rel=T2relmaxαS2σα=σ0+σ1=3\nu^{+rel}=T^{rel}_{2}\leq\max_{\alpha\in S_{2}}\sigma\cdot\alpha=\sigma_{0}+\sigma_{1}=3

Let xx be the number of 1’s in σ\sigma. Then

ν+rel\displaystyle\nu^{+rel} 12+12(rp|σ|1)\displaystyle\geq-\dfrac{1}{2}+\dfrac{1}{2}(rp-|\sigma|_{1})
=12+12(r(4+x)(2+x))\displaystyle=-\dfrac{1}{2}+\dfrac{1}{2}\left(r(4+x)-(2+x)\right)
=2r1+(r1)x12\displaystyle=2r-1+\dfrac{(r-1)x-1}{2}

Hence, we must have r=2r=2 and x=1x=1. ∎

Theorem 1.4 comes as a corollary.

Theorem 1.4.

Let KK be a knot in S3S^{3} with ν+(K)=0\nu^{+}(K)=0 and let mm be a positive integer. If the mm-surgery on KK is lensbordant, then it is smoothly rational homology cobordant to L(m,m1)L(m,m-1).

Proof of Theorem 1.4.

When ν+=0\nu^{+}=0, Lemma 4.4 implies that either r=1r=1, or r=2r=2 with σ=(1)\sigma=(1).

The changemaker σ=(1)\sigma=(1) represents S3S^{3}. Hence, in the r=2r=2 with σ=(1)\sigma=(1) case, m=4m=4 and YY is smoothly rational homology cobordant to S3S^{3}. Since S3S^{3} is smoothly rational homology cobordant to L(4,3)L(4,3), YY is also smoothly rational homology cobordant to L(4,3)L(4,3).

When r=1r=1, the conditions in Lemma 4.2 become the conditions in [12, Lemma 2.10]. Therefore, when ν+(K)=0\nu^{+}(K)=0, all entries of σ\sigma are 1’s [12, Lemma 2.12]. These changemakers correspond to L(m,m1)L(m,m-1). ∎

Given Lemma 4.3 and 4.4, from this point onwards, until the end of this section, we assume that r2r\geq 2 and V0rel2V^{rel}_{0}\geq 2.

Now we can define

μ:=min1i<V0rel(TirelTi1rel)\mu:=\min_{1\leq i<V^{rel}_{0}}(T^{rel}_{i}-T^{rel}_{i-1})

Note that with r2r\geq 2, μ\mu can be 0 here. This is very different from what McCoy did in [12] where μ\mu must be at least 1.

Lemma 4.5.
V0rel{18((r2p2r+1)(n+1)) if r is even and p is odd18(r2p(n+1)) otherwiseV^{rel}_{0}\geq\begin{cases}\frac{1}{8}((r^{2}p-2r+1)-(n+1))\text{ if }r\text{ is even and }p\text{ is odd}\\ \frac{1}{8}(r^{2}p-(n+1))\text{ otherwise}\end{cases}
Proof.

At the start of the proof of Lemma 4.2, we see that

8V0rel=𝔠2+(n+1)-8V^{rel}_{0}=\mathfrak{c}^{2}+(n+1)

for some characteristic vector 𝔠\mathfrak{c} satisfying

𝔠,σ+rp={1 if r is even and p is odd0 otherwise\langle\mathfrak{c},\sigma\rangle+rp=\begin{cases}1\text{ if }r\text{ is even and }p\text{ is odd}\\ 0\text{ otherwise}\end{cases}

By Cauchy-Schwarz inequality and the fact that σ2=p\sigma^{2}=-p, we have

|𝔠2|=|𝔠2||σ2|p|𝔠,σ|2p{(rp1)2p if r is even and p is odd(rp)2p otherwise|\mathfrak{c}^{2}|=\frac{|\mathfrak{c}^{2}||\sigma^{2}|}{p}\geq\frac{|\langle\mathfrak{c},\sigma\rangle|^{2}}{p}\geq\begin{cases}\frac{(rp-1)^{2}}{p}\text{ if }r\text{ is even and }p\text{ is odd}\\ \frac{(rp)^{2}}{p}\text{ otherwise}\end{cases}

Simplifying the expressions and considering the fact that |𝔠2||\mathfrak{c}^{2}|\in\mathbb{Z}, we have

|𝔠2|{r2p2r+1 if r is even and p is oddr2p otherwise|\mathfrak{c}^{2}|\geq\begin{cases}r^{2}p-2r+1\text{ if }r\text{ is even and }p\text{ is odd}\\ r^{2}p\text{ otherwise}\end{cases}

Result follows by substituting this back to 8V0rel=𝔠2+(n+1)-8V^{rel}_{0}=\mathfrak{c}^{2}+(n+1). ∎

Lemma 4.6.

If μ2\mu\geq 2, then V0rel=2V_{0}^{rel}=2 and σ0=2\sigma_{0}=2 and σ1=1\sigma_{1}=1.

Proof.

Suppose μ2\mu\geq 2. We define

M:={18((r2p2r+1)(n+1))1 if r is even and p is odd18(r2p(n+1))1 otherwiseM:=\begin{cases}\lceil\frac{1}{8}((r^{2}p-2r+1)-(n+1))\rceil-1\text{ if }r\text{ is even and }p\text{ is odd}\\ \lceil\frac{1}{8}(r^{2}p-(n+1))\rceil-1\text{ otherwise}\end{cases}

which is an analogue of N1N-1 in [12, §2].

By Lemma 4.5, we have V0rel>MV_{0}^{rel}>M. Hence, by Lemma 4.2, we know that

TMrel=maxαSMσαT^{rel}_{M}=\displaystyle\max_{\alpha\in S_{\leq M}}\sigma\cdot\alpha

Take such an α\alpha.

First of all, we argue that αSM\alpha\in S_{M} (without the \leq). If αSM\alpha\not\in S_{M}, then we must have M1M\geq 1 and maxαSMσα=maxαSM1σα\max\limits_{\alpha\in S_{\leq M}}\sigma\cdot\alpha=\max\limits_{\alpha\in S_{\leq M-1}}\sigma\cdot\alpha. This implies TMrel=TM1relT^{rel}_{M}=T^{rel}_{M-1}, which contradicts with μ2\mu\geq 2. Hence, αSM\alpha\in S_{M}.

For any index ll with αl>0\alpha_{l}>0, if we consider α\alpha^{{}^{\prime}} defined with αi={αi if ilαi1 if i=l\alpha^{{}^{\prime}}_{i}=\begin{cases}\alpha_{i}\text{ if }i\neq l\\ \alpha_{i}-1\text{ if }i=l\end{cases}, we have αSMαl\alpha^{{}^{\prime}}\in S_{\leq M-\alpha_{l}}. So we have

TMαlrelασ=ασσl=TMrelσlT^{rel}_{M-\alpha_{l}}\geq\alpha^{{}^{\prime}}\cdot\sigma=\alpha\cdot\sigma-\sigma_{l}=T^{rel}_{M}-\sigma_{l}

Hence,

σlTMrelTMαlrelαlμ\sigma_{l}\geq T^{rel}_{M}-T^{rel}_{M-\alpha_{l}}\geq\alpha_{l}\mu

Since μ2\mu\geq 2, from our discussion above, for all index ll we have σl2αl\sigma_{l}\geq 2\alpha_{l} (this also holds when αl=0\alpha_{l}=0 because the right hand side becomes 0). Hence,

M\displaystyle M =12lαl(αl+1)\displaystyle=\dfrac{1}{2}\displaystyle\sum_{l}\alpha_{l}(\alpha_{l}+1)
12lσl2(σl2+1)\displaystyle\leq\dfrac{1}{2}\displaystyle\sum_{l}\dfrac{\sigma_{l}}{2}\left(\dfrac{\sigma_{l}}{2}+1\right)
=18lσl(σl+2)\displaystyle=\dfrac{1}{8}\displaystyle\sum_{l}\sigma_{l}(\sigma_{l}+2)
=18(p+2|σ|1)\displaystyle=\dfrac{1}{8}(p+2|\sigma|_{1})

Combining this with the definition of MM, we have

18(p+2|σ|1){18((r2p2r+1)(n+1))1 if r is even and p is odd18(r2p(n+1))1 otherwise\dfrac{1}{8}(p+2|\sigma|_{1})\geq\begin{cases}\left\lceil\frac{1}{8}((r^{2}p-2r+1)-(n+1))\right\rceil-1\text{ if }r\text{ is even and }p\text{ is odd}\\ \left\lceil\frac{1}{8}(r^{2}p-(n+1))\right\rceil-1\text{ otherwise}\end{cases}

Since r2r\geq 2 and |σ|1(n+1)|\sigma|_{1}\geq(n+1) (because σ\sigma has no 0 entries), we have

18(p+2|σ|1){18((4p4+1)|σ|1)1 if r is even and p is odd18(4p|σ|1)1 otherwise\dfrac{1}{8}(p+2|\sigma|_{1})\geq\begin{cases}\frac{1}{8}((4p-4+1)-|\sigma|_{1})-1\text{ if }r\text{ is even and }p\text{ is odd}\\ \frac{1}{8}(4p-|\sigma|_{1})-1\text{ otherwise}\end{cases}

Rearranging this, we get

{113p|σ|1 if r is even and p is odd83p|σ|1 otherwise\begin{cases}\dfrac{11}{3}\geq p-|\sigma|_{1}\text{ if }r\text{ is even and }p\text{ is odd}\\[12.91663pt] \dfrac{8}{3}\geq p-|\sigma|_{1}\text{ otherwise}\end{cases}

Since p|σ|1p-|\sigma|_{1} is an integer, this implies

3l(σl2σl)3\geq\displaystyle\sum_{l}(\sigma_{l}^{2}-\sigma_{l})

Therefore, the entries of σ\sigma is either a 2 followed by a bunch of 1’s or simply all 1’s.

Since V0rel2V^{rel}_{0}\geq 2, we have T0rel=0T^{rel}_{0}=0 and T1rel=σ0T^{rel}_{1}=\sigma_{0}. Also, if V0rel3V^{rel}_{0}\geq 3, we have T2rel=σ0+σ1T^{rel}_{2}=\sigma_{0}+\sigma_{1}. Therefore, the only way for μ\mu to be at least 2 is when V0rel=2V^{rel}_{0}=2 and σ\sigma is a 2 followed by a bunch of 1’s. ∎

Now we can prove Theorem 1.7. Statement (1) of Theorem 1.7 has already been proven in Lemma 4.3. So we only need to prove statements (2) and (3) of Theorem 1.7.

Proof of statement (2) and (3) of Theorem 1.7.

Fix the knot KK. Fix r2r\geq 2. Also, if rr is even, we fix the parity of pp (odd or even).

We want to show that these together determines pp, which determines the surgery slope r2pr^{2}p since rr is fixed.

These data determines the relevant coefficients. So, it suffices to show that the relevant coefficients together with the value of rr determines σ\sigma (which determines pp).

The algorithm McCoy described in the proof of [12, Lemma 2.16] shows that statement (1) of our Lemma 4.2 determines all entries of σ\sigma strictly greater than μ\mu. Together with Lemma 4.6, this implies that all entries of σ\sigma except for the 1’s are determined, except for a special case which was addressed in Lemma 4.4. Given that r2r\geq 2, statement (2) of Lemma 4.2 determines the number of 1’s. ∎

[Uncaptioned image]

Figure 2. Flowchart showing McCoy’s algorithm and argument, but in notations used in this paper.

5. Some bounds on the surgery slope r2pr^{2}p

In this section, we establish some inequalities involving the surgery slope r2pr^{2}p.

Theorem 1.8.

Suppose there is an (r,p)(r,p)-lensbordant surgery on a knot KK in S3S^{3} with associated changemaker σ\sigma, then

2ν+(K)+2(r1)r2pr|σ|12ν+(K)2\nu^{+}(K)+2(r-1)\geq r^{2}p-r|\sigma|_{1}\geq 2\nu^{+}(K)
Proof.

When ν+rel=0\nu^{+rel}=0, Lemma 4.4 implies that either r=2r=2 with σ=(1)\sigma=(1), or r=1r=1. When r=2r=2 and σ=(1)\sigma=(1), the inequality becomes 2202\geq 2\geq 0. When r=1r=1, V0=V0rel=0V_{0}=V^{rel}_{0}=0, and hence ν+=0\nu^{+}=0. As discussed in the proof of Theorem 1.4, σ\sigma must be (1,,1)(1,\dots,1) in this case, and the inequality becomes 0000\geq 0\geq 0.

Now we assume that ν+rel0\nu^{+rel}\neq 0

Case 1: rr is odd or pp is even
We have ν+rel=12(rp|σ|1)\nu^{+rel}=\dfrac{1}{2}(rp-|\sigma|_{1}). Hence,

V12(rp|σ|1)rel=0andV12(rp|σ|1)1rel0V^{rel}_{\tfrac{1}{2}(rp-|\sigma|_{1})}=0\quad\text{and}\quad V^{rel}_{\tfrac{1}{2}(rp-|\sigma|_{1})-1}\neq 0

So we have

V12(r2pr|σ|1)=0andV12(r2pr|σ|1)r0V_{\tfrac{1}{2}(r^{2}p-r|\sigma|_{1})}=0\quad\text{and}\quad V_{\tfrac{1}{2}(r^{2}p-r|\sigma|_{1})-r}\neq 0

This implies

12(r2pr|σ|1)ν+>12(r2pr|σ|1)r\dfrac{1}{2}(r^{2}p-r|\sigma|_{1})\geq\nu^{+}>\dfrac{1}{2}(r^{2}p-r|\sigma|_{1})-r

Rearranging, we have

2ν++2r>r2pr|σ|12ν+2\nu^{+}+2r>r^{2}p-r|\sigma|_{1}\geq 2\nu^{+}

Since p|σ|1p\equiv|\sigma|_{1} (mod 2), the value r2pr|σ|1r^{2}p-r|\sigma|_{1} must be an even integer. Hence

2ν++2(r1)r2pr|σ|12ν+2\nu^{+}+2(r-1)\geq r^{2}p-r|\sigma|_{1}\geq 2\nu^{+}

Case 2: rr is even and pp is odd
We have ν+rel=12+12(rp|σ|1)\nu^{+rel}=-\dfrac{1}{2}+\dfrac{1}{2}(rp-|\sigma|_{1}). Hence,

V12+12(rp|σ|1)rel=0andV32+12(rp|σ|1)rel0V^{rel}_{-\tfrac{1}{2}+\tfrac{1}{2}(rp-|\sigma|_{1})}=0\quad\text{and}\quad V^{rel}_{-\tfrac{3}{2}+\tfrac{1}{2}(rp-|\sigma|_{1})}\neq 0

So we have

V12(r2pr|σ|1)=0andV12(r2pr|σ|1)r0V_{\tfrac{1}{2}(r^{2}p-r|\sigma|_{1})}=0\quad\text{and}\quad V_{\tfrac{1}{2}(r^{2}p-r|\sigma|_{1})-r}\neq 0

The rest of the proof is the same as Case 1. ∎

Now we prove Corollary 1.9.

Corollary 1.9.

Suppose there is an (r,p)(r,p)-lensbordant surgery on a knot KK in S3S^{3} with associated changemaker σ\sigma, then
(1) If r3r\geq 3, then 2r2(r1)(r2)ν+(K)r2p2ν+(K)+r\dfrac{2r^{2}}{(r-1)(r-2)}\nu^{+}(K)\geq r^{2}p\geq 2\nu^{+}(K)+r
(2) If r2r\geq 2, then 4ν+(K)+5r2p2ν+(K)+r4\nu^{+}(K)+5\geq r^{2}p\geq 2\nu^{+}(K)+r

Proof.

Since |σ|11|\sigma|_{1}\geq 1, Theorem 1.8 implies that r2p2ν++rr^{2}p\geq 2\nu^{+}+r.

Since p|σ|1p\geq|\sigma|_{1}, Theorem 1.8 implies that 2ν++2(r1)r(r1)p2\nu^{+}+2(r-1)\geq r(r-1)p. So we have

(8) 2rr1ν++2rr2p\dfrac{2r}{r-1}\nu^{+}+2r\geq r^{2}p

Then we have

r2rr2p=r2p2rr2p(2rr1ν++2r)2r=2rr1ν+\dfrac{r-2}{r}r^{2}p=r^{2}p-\dfrac{2}{r}r^{2}p\leq\left(\dfrac{2r}{r-1}\nu^{+}+2r\right)-2r=\dfrac{2r}{r-1}\nu^{+}

Statement (1) follows.

When r6r\geq 6, 2r2(r1)(r2)<4\dfrac{2r^{2}}{(r-1)(r-2)}<4. So, we only need to check statement (2) for r=2,3,4,5r=2,3,4,5.

When r=2r=2, (8) implies r2p4ν++4r^{2}p\leq 4\nu^{+}+4

By Lemma 4.4, when r3r\geq 3, we have ν+1\nu^{+}\geq 1.
When r=3r=3, (8) implies r2p3ν++64ν++5r^{2}p\leq 3\nu^{+}+6\leq 4\nu^{+}+5.

When r=4r=4 and ν+3\nu^{+}\geq 3, (8) implies r2p83ν++8=4ν++443(ν+3)4ν++4r^{2}p\leq\tfrac{8}{3}\nu^{+}+8=4\nu^{+}+4-\tfrac{4}{3}(\nu^{+}-3)\leq 4\nu^{+}+4.
When r=4r=4 and ν+2\nu^{+}\leq 2, statement (1) implies r2p163ν+=4ν++8343(2ν+)4ν++83r^{2}p\leq\tfrac{16}{3}\nu^{+}=4\nu^{+}+\tfrac{8}{3}-\tfrac{4}{3}(2-\nu^{+})\leq 4\nu^{+}+\tfrac{8}{3}.

When r=5r=5 and ν+4\nu^{+}\geq 4, (8) implies r2p52ν++10=4ν++432(ν+4)4ν++4r^{2}p\leq\tfrac{5}{2}\nu^{+}+10=4\nu^{+}+4-\tfrac{3}{2}(\nu^{+}-4)\leq 4\nu^{+}+4.
When r=5r=5 and ν+3\nu^{+}\leq 3, statement (1) implies r2p256ν+=4ν++1216(3ν+)4ν++12r^{2}p\leq\tfrac{25}{6}\nu^{+}=4\nu^{+}+\tfrac{1}{2}-\tfrac{1}{6}(3-\nu^{+})\leq 4\nu^{+}+\tfrac{1}{2}. ∎

6. Structure of relevant coefficients

In the introduction, the remark after the statement of Theorem 1.10 states that the surgery slope r2pr^{2}p is very close to 8V0(K)8V_{0}(K). The precise statement is

Theorem 6.1.

(1) If r,pr,p are both even, then r2p=8V0(K)r^{2}p=8V_{0}(K).
(2) If rr is even and pp is odd, then 8V0(K)2rr2p8V0(K)+2r8V_{0}(K)-2r\leq r^{2}p\leq 8V_{0}(K)+2r.
(3) If rr is odd, then 8V0(K)3odd(σ)r2p8V0(K)+odd(σ)8V_{0}(K)-3odd(\sigma)\leq r^{2}p\leq 8V_{0}(K)+odd(\sigma)
where odd(σ)odd(\sigma) is the number of odd entries in σ\sigma.

Furthermore, in the case when rr is odd, r2p=8V0(K)+odd(σ)r^{2}p=8V_{0}(K)+odd(\sigma) if and only if the even entries of σ\sigma can be partitioned into two sets with equal sum.

Note that Theorem 1.10 is statement (1) of Theorem 6.1. The aim of this section is to gain a deeper understanding on relevant coefficients, and then prove Theorem 6.1.


rr even rr odd
pp even r2p=8V0r^{2}p=8V_{0} 8V03odd(σ)r2p8V0+odd(σ)8V_{0}-3odd(\sigma)\leq r^{2}p\leq 8V_{0}+odd(\sigma)
upper bound is equality if and only if
pp odd 8V02rr2p8V0+2r8V_{0}-2r\leq r^{2}p\leq 8V_{0}+2r the even entries of σ\sigma can be partitioned
into two sets with equal sum

Table 1. Table describing Theorem 6.1.


The TmrelT^{rel}_{m} in statement (1) of Lemma 4.2 is defined with the relevant coefficients. Besides mm, it also depends on the knot KK, the value of rr, and the parity of pp if rr is even. However, the right hand side maxαSmσα\displaystyle\max_{\alpha\in S_{\leq m}}\sigma\cdot\alpha only depends on mm and σ\sigma. Hence, it makes sense to think of that as some coefficients coming from σ\sigma itself.

Definition 6.2.

For m0m\geq 0, we define

Tmσ:=maxαSmσαT^{\sigma}_{m}:=\displaystyle\max_{\alpha\in S_{\leq m}}\sigma\cdot\alpha

and we define the relevant coefficients ViσV^{\sigma}_{i} (for i>0i>0) of the changemaker vector σ\sigma to be the monotonic increasing sequence of positive integers such that TmσT^{\sigma}_{m} is the number of relevant coefficients less than or equal to mm. Equivalently,

Viσ:=min{miTmσ},i>0V^{\sigma}_{i}:=\min\{m\mid i\leq T^{\sigma}_{m}\},\quad i\in\mathbb{Z}_{>0}

We also define V0σV^{\sigma}_{0} to be 0, but we do not call it a relevant coefficient to avoid making the above definition more complicated.

Now statement (1) and (3) of Lemma 4.2 can be rewritten as:

Tmrel=Tmσ for all m<V0relT^{rel}_{m}=T^{\sigma}_{m}\text{ for all }m<V^{rel}_{0}

and

TV0relrelTV0relσT^{rel}_{V^{rel}_{0}}\leq T^{\sigma}_{V^{rel}_{0}}

Together they imply

Virel=Vν+reliσ for all 0i<ν+relV^{rel}_{i}=V^{\sigma}_{\nu^{+rel}-i}\text{ for all }0\leq i<\nu^{+rel}

Recall that the definition of SmS_{\leq m} is

Sm:={α0dim(σ)Σαj(αj+1)2m}S_{\leq m}:=\{\alpha\in\mathbb{Z}^{dim(\sigma)}_{\geq 0}\mid\Sigma\alpha_{j}(\alpha_{j}+1)\leq 2m\}

Notice that 12αj(αj+1)\tfrac{1}{2}\alpha_{j}(\alpha_{j}+1) is exactly 1+2++αj1+2+\dots+\alpha_{j}. Hence, we can interpret TmσT^{\sigma}_{m} as the answer to the following combinatorial problem:

Consider the following game. There is a shop where you can spend coins to buy items. The available items are labeled I0,,InI_{0},\dots,I_{n}. You can buy any non-negative integer amount of each of the items. For each individual item, it costs 1 coin to buy the first one, 2 coins to buy the second one, 3 coins to buy the third one, etc. After finish doing all the purchases, each IiI_{i} you own gives you σi\sigma_{i} points. If you start with mm coins, what is the highest amount of points that you can get?

A way to approach this game is to consider the coin cost per point in each purchase option. Purchasing the item IiI_{i} for the kthk^{th} time costs kk coins to obtain σi\sigma_{i} points, hence it is a kσi\dfrac{k}{\sigma_{i}} coin per point purchase. One intuitive thing to do is the use a greedy algorithm by always making the purchase with the lowest coin cost per point among all available purchase options until running out of coins.

This greedy algorithm does not always give the optimal solution. At the end, you might have some coins left that is not enough to purchase the next lowest coin per point purchase option and are forced to purchase a cheaper option with a higher cost per point, or simply having some leftover coins with no available purchase options.

However, consider a modified version of the game where you are allowed to make fractional purchases under the same coin per point. For example, being able to buy (4+23)(4+\tfrac{2}{3}) copies of IiI_{i} by spending (1+2+3+4+23×5)(1+2+3+4+\tfrac{2}{3}\times 5) coins, gaining (4+23)σi(4+\tfrac{2}{3})\sigma_{i} points.

In this modified version of the game, a greedy algorithm will not result in any leftover coins, and all the purchased options have a lower or equal coin per point than any unpurchased option. Hence, a greedy algorithm must give the optimal solution.

Definition 6.3.

We call a greedy algorithm on the modified version of the game a rational greedy algorithm. We write Tmσ,T^{\sigma,\mathbb{Q}}_{m} to denote the amount of points obtained with mm coins under the rational greedy algorithm.

Given m,σm,\sigma, we say that the resulting rational greedy algorithm is up to xx\in\mathbb{Q} coin per point if the resulting rational greedy algorithm makes a non-zero a amount of purchase options with xx coin per point and no purchase options with >x>x coin per point.

Because of the way rational greedy algorithm works, a rational greedy algorithm up to xx coin per point must make all available purchases that are <x<x coin per point.

Also, given σ\sigma and mm, if a rational greedy algorithm is up to xx coin per point, then all other optimal solutions to the modified version of the game must also take the same form: Purchasing all available purchases that are <x<x coin per point, and not purchasing any options that are >x>x coin per point. That is because any other way of spending all mm coins will result in having a higher average coin per point, and thus a lower total point at the end.

Hence, in this modified version of the game, all optimal solutions must be rational greedy algorithms up to xx coin per cost. The only difference allowed in different optimal solutions is picking different xx coin per point purchase options.

Definition 6.4.

If an optimal solution in the modified version of the game does not include fractional purchases, we say that the solution is integral. Equivalently, an integral optimal solution in the modified version of the game is a solution that is valid in the original version of the game.

If an optimal solution in the modified version of the game is a rational greedy algorithm up to xx coin per point that purchases all options that are xx coin per point, we say that the solution is full.

Clearly, a full rational greedy algorithm must also be integral.

All purchase options in the original version of the game are valid purchase options in the modified version of the game. Hence, Tmσ,TmσT^{\sigma,\mathbb{Q}}_{m}\geq T^{\sigma}_{m}. In particular, Tmσ,=TmσT^{\sigma,\mathbb{Q}}_{m}=T^{\sigma}_{m} if and only if there exist an integral implementation of the rational greedy algorithm to the modified version of the game.

In this section we will use these ideas to help investigating the structure of relevant coefficients.

Lemma 6.5.

For any positive integer xx,

Tx2p+x|σ|12σ=xpT^{\sigma}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}}=xp
Proof.

Consider the full rational greedy algorithm up to xx coin per point. So, for each ii, it purchases xσix\sigma_{i} copies of the item IiI_{i}. The total amount of coins spent is

12i(xσi)(xσi+1)\displaystyle\dfrac{1}{2}\sum\limits_{i}(x\sigma_{i})(x\sigma_{i}+1) =12x2iσi+12xiσi\displaystyle=\dfrac{1}{2}x^{2}\sum\limits_{i}\sigma_{i}+\dfrac{1}{2}x\sum\limits_{i}\sigma_{i}
=x2p+x|σ|12\displaystyle=\dfrac{x^{2}p+x|\sigma|_{1}}{2}

And the number of points gained is

i(xσi)σi=xp\sum\limits_{i}(x\sigma_{i})\sigma_{i}=xp

Hence, Tx2p+x|σ|12σ,=xpT^{\sigma,\mathbb{Q}}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}}=xp. Since the greedy algorithm is full, it must be integral. Therefore,

Tx2p+x|σ|12σ,=Tx2p+x|σ|12σT^{\sigma,\mathbb{Q}}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}}=T^{\sigma}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}}

Corollary 6.6.

For any positive integer xx,

Vxpσ=x2p+x|σ|12V^{\sigma}_{xp}=\dfrac{x^{2}p+x|\sigma|_{1}}{2}
Proof.

From Lemma 6.5, we know there is exactly xpxp relevant coefficients less than or equal to x2p+x|σ|12\dfrac{x^{2}p+x|\sigma|_{1}}{2}. Since the relevant coefficients of a changemaker vector are monotonic increasing, we know that

Vxpσx2p+x|σ|12V^{\sigma}_{xp}\leq\dfrac{x^{2}p+x|\sigma|_{1}}{2}

To prove Vxpσ=x2p+x|σ|12V^{\sigma}_{xp}=\dfrac{x^{2}p+x|\sigma|_{1}}{2}, it suffices to show that

Tx2p+x|σ|121σ<xpT^{\sigma}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}-1}<xp

i.e. there are strictly less than xpxp relevant coefficients that are less than or equal to x2p+x|σ|121\dfrac{x^{2}p+x|\sigma|_{1}}{2}-1 (which implies that VxpσV^{\sigma}_{xp} cannot be less than or equal to x2p+x|σ|121\tfrac{x^{2}p+x|\sigma|_{1}}{2}-1)

This follows from the way Lemma 6.5 is proved. xpxp points is obtained from a rational greedy algorithm with x2p+x|σ|12\tfrac{x^{2}p+x|\sigma|_{1}}{2} coins. Since every coin contributes to some points under a rational greedy algorithm, strictly less than xpxp points is obtained from a rational greedy algorithm with strictly less than x2p+x|σ|12\tfrac{x^{2}p+x|\sigma|_{1}}{2} coins. Hence

Tx2p+x|σ|121σTx2p+x|σ|121σ,<xpT^{\sigma}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}-1}\leq T^{\sigma,\mathbb{Q}}_{\tfrac{x^{2}p+x|\sigma|_{1}}{2}-1}<xp

Roughly speaking, Corollary 6.6 reveals that the relevant coefficents increase in a quadratic rate. The next thing we aim to show is that the first pp relevant coefficients of σ\sigma determines the rest of them in a simple manner. Precisely, we want to show that for all positive integer xx and 0k<p0\leq k<p,

(9) Vxp+kσ=x2p+x|σ|12+xk+VkσV^{\sigma}_{xp+k}=\dfrac{x^{2}p+x|\sigma|_{1}}{2}+xk+V^{\sigma}_{k}

This is particularly useful for those who wish to do some programming on this topic because the size of SmS_{\leq m} increases quickly with mm, which can cause run-time and memory issues. Based on data from some programming efforts related to this paper, when m=100m=100, there are 157,452 α>0something\alpha\in\mathbb{Z}^{something}_{>0} such that α0α1\alpha_{0}\geq\alpha_{1}\geq\dots and jαj(αj+1)2m\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\leq 2m. When m=200m=200, the number becomes 13,552,451.

To help proving (9), we prove some more lemmas first.

Lemma 6.7.

For any positive integer xx,

Tx2px|σ|12σ=xp|σ|1T^{\sigma}_{\tfrac{x^{2}p-x|\sigma|_{1}}{2}}=xp-|\sigma|_{1}
Proof.

Consider the full rational greedy algorithm that purchases all options strictly less than xx coin per point. So, for each ii, it purchases xσi1x\sigma_{i}-1 copies of the item IiI_{i}. The total amount of coins spent is

12i(xσi1)(xσi)=x2px|σ|12\dfrac{1}{2}\sum\limits_{i}(x\sigma_{i}-1)(x\sigma_{i})=\dfrac{x^{2}p-x|\sigma|_{1}}{2}

And the number of points gained is

i(xσi1)σi=xp|σ|1\sum\limits_{i}(x\sigma_{i}-1)\sigma_{i}=xp-|\sigma|_{1}

Result follows from the fact that the rational greedy algorithm is full, and therefore integral. ∎

Corollary 6.8.
Vxp|σ|1σ=x2px|σ|12V^{\sigma}_{xp-|\sigma|_{1}}=\dfrac{x^{2}p-x|\sigma|_{1}}{2}
Proof.

Since Lemma 6.7 came from an integral rational greedy algorithm, the argument used in the proof of Corollary 6.6 can be applied. ∎

Lemma 6.9.

For any positive integer xx, if ax2p+x|σ|12a\leq\dfrac{x^{2}p+x|\sigma|_{1}}{2} and Ta1σ<TaσT^{\sigma}_{a-1}<T^{\sigma}_{a} and Tbσ=Ta1σT^{\sigma}_{b}=T^{\sigma}_{a-1}, then abxa-b\leq x.

Proof.

Let αSb\alpha\in S_{\leq b} be such that ασ=Tbσ\alpha\cdot\sigma=T^{\sigma}_{b}.

To show that abxa-b\leq x, it suffices to show that αSb+x\exists\alpha^{\prime}\in S_{\leq b+x} such that ασTbσ+1\alpha^{\prime}\cdot\sigma\geq T^{\sigma}_{b}+1 because the existence of such an α\alpha^{\prime} implies Tb+xσTbσ+1>Tbσ=Ta1σT^{\sigma}_{\leq b+x}\geq T^{\sigma}_{b}+1>T^{\sigma}_{b}=T^{\sigma}_{a-1}, which implies b+x>a1b+x>a-1, which implies abxa-b\leq x.

Let ii be minimal such that αi<xσi\alpha_{i}<x\sigma_{i}. Such an ii much exists, otherwise

12jαj(αj+1)x2p+x|σ|12a>b\dfrac{1}{2}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)\geq\dfrac{x^{2}p+x|\sigma|_{1}}{2}\geq a>b

which contradicts αSb\alpha\in S_{\leq b}.

Since σ\sigma is a changemaker, σi1,,σik\exists\sigma_{i_{1}},\dots,\sigma_{i_{k}} such that σi1++σik=σi1\sigma_{i_{1}}+\dots+\sigma_{i_{k}}=\sigma_{i}-1. (when σi=1\sigma_{i}=1, we set k=0k=0 and this is an empty sum)

Define α\alpha^{\prime} be such that

αj={αj+1 if j=iαj1 if j=i1,,ikαj otherwise\alpha^{\prime}_{j}=\begin{cases}\alpha_{j}+1\text{ if }j=i\\ \alpha_{j}-1\text{ if }j=i_{1},\dots,i_{k}\\ \alpha_{j}\text{ otherwise}\end{cases}

Then

ασ\displaystyle\alpha^{\prime}\cdot\sigma =ασ+σiσi1σik\displaystyle=\alpha\cdot\sigma+\sigma_{i}-\sigma_{i_{1}}-\dots-\sigma_{i_{k}}
=ασ+1\displaystyle=\alpha\cdot\sigma+1
=Tbσ+1\displaystyle=T^{\sigma}_{b}+1

and also

αSb+(αi+1)αi1αik\alpha^{\prime}\in S_{\leq b+(\alpha_{i}+1)-\alpha_{i_{1}}-\dots-\alpha_{i_{k}}}

We complete the proof by noting

b+(αi+1)αi1αik\displaystyle b+(\alpha_{i}+1)-\alpha_{i_{1}}-\dots-\alpha_{i_{k}} <b+(nσi+1)nσi1nσik\displaystyle<b+(n\sigma_{i}+1)-n\sigma_{i_{1}}-\dots-n\sigma_{i_{k}}
=b+x+1\displaystyle=b+x+1

Hence

b+(αi+1)αi1αikb+xb+(\alpha_{i}+1)-\alpha_{i_{1}}-\dots-\alpha_{i_{k}}\leq b+x

Corollary 6.10.

For any positive integer xx, if 0<axp0<a\leq xp, then VaσVa1σxV^{\sigma}_{a}-V^{\sigma}_{a-1}\leq x

Proof.

If Vaσ=Va1σV^{\sigma}_{a}=V^{\sigma}_{a-1}, then the result is trivial. So, we assume Vaσ>Va1σV^{\sigma}_{a}>V^{\sigma}_{a-1}.

By Corollary 6.6, we know that Vax2p+x|σ|12V_{a}\leq\tfrac{x^{2}p+x|\sigma|_{1}}{2}. Since there is at least one relevant coefficient with value VaσV^{\sigma}_{a} (namely VaσV^{\sigma}_{a} itself), we know that TVa1σ<TVaσT^{\sigma}_{V_{a}-1}<T^{\sigma}_{V_{a}}. Also, since Vaσ>Va1σV^{\sigma}_{a}>V^{\sigma}_{a-1}, we have TVa1σ=TVa1σ=a1T^{\sigma}_{V_{a-1}}=T^{\sigma}_{V_{a}-1}=a-1. Result follows from Lemma 6.9. ∎

Lemma 6.11.

For any positive integer xx, if a>x2px|σ|12a>\dfrac{x^{2}p-x|\sigma|_{1}}{2} and Taσ>Ta1σT^{\sigma}_{a}>T^{\sigma}_{a-1} and TaσTbσ2T^{\sigma}_{a}-T^{\sigma}_{b}\geq 2, then abx+1a-b\geq x+1.

Proof.

Let αSa\alpha\in S_{\leq a} be such that ασ=Taσ\alpha\cdot\sigma=T^{\sigma}_{a}.

To show that abx+1a-b\geq x+1, it suffices to show that αSax\exists\alpha^{\prime}\in S_{\leq a-x} such that ασTaσ1\alpha^{\prime}\cdot\sigma\geq T^{\sigma}_{a}-1 because the existence of such an α\alpha^{\prime} implies TaxσTaσ1Tbσ+1>TbσT^{\sigma}_{\leq a-x}\geq T^{\sigma}_{a}-1\geq T^{\sigma}_{b}+1>T^{\sigma}_{b}, which implies ax>ba-x>b, which implies abx+1a-b\geq x+1.

Let ii be minimal such that αixσi\alpha_{i}\geq x\sigma_{i}. Such an ii much exists, otherwise

12jαj(αj+1)\displaystyle\dfrac{1}{2}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1) 12j(xσj1)(xσj)\displaystyle\leq\dfrac{1}{2}\sum\limits_{j}(x\sigma_{j}-1)(x\sigma_{j})
=x2px|σ|12\displaystyle=\dfrac{x^{2}p-x|\sigma|_{1}}{2}
<a\displaystyle<a

which implies αSa1\alpha\in S_{\leq a-1}, which contradicts with Taσ>Ta1σT^{\sigma}_{a}>T^{\sigma}_{a-1}.

Since σ\sigma is a changemaker, σi1,,σik\exists\sigma_{i_{1}},\dots,\sigma_{i_{k}} such that σi1++σik=σi1\sigma_{i_{1}}+\dots+\sigma_{i_{k}}=\sigma_{i}-1. (when σi=1\sigma_{i}=1, we set k=0k=0 and this is an empty sum)

Define α\alpha^{\prime} be such that

αj={αj1 if j=iαj+1 if j=i1,,ikαj otherwise\alpha^{\prime}_{j}=\begin{cases}\alpha_{j}-1\text{ if }j=i\\ \alpha_{j}+1\text{ if }j=i_{1},\dots,i_{k}\\ \alpha_{j}\text{ otherwise}\end{cases}

Then

ασ\displaystyle\alpha^{\prime}\cdot\sigma =ασσi+σi1++σik\displaystyle=\alpha\cdot\sigma-\sigma_{i}+\sigma_{i_{1}}+\dots+\sigma_{i_{k}}
=ασ1\displaystyle=\alpha\cdot\sigma-1
=Taσ1\displaystyle=T^{\sigma}_{a}-1

and also

αSaαi+(αi1+1)++(αik+1)\alpha^{\prime}\in S_{\leq a-\alpha_{i}+(\alpha_{i_{1}}+1)+\dots+(\alpha_{i_{k}}+1)}

We complete the proof by noting

aαi+(αi1+1)++(αik+1)\displaystyle a-\alpha_{i}+(\alpha_{i_{1}}+1)+\dots+(\alpha_{i_{k}}+1) aαi+nσi1++nσik\displaystyle\leq a-\alpha_{i}+n\sigma_{i_{1}}+\dots+n\sigma_{i_{k}}
=an(αinσi)\displaystyle=a-n-(\alpha_{i}-n\sigma_{i})
an\displaystyle\leq a-n

Corollary 6.12.

For any positive integer xx, if axp|σ|1a\geq xp-|\sigma|_{1}, then Va+1σVaσxV^{\sigma}_{a+1}-V^{\sigma}_{a}\geq x

Proof.

By Lemma 6.7, there are only xp|σ|1xp-|\sigma|_{1} relevant coefficients less than or equal to x2px|σ|12\tfrac{x^{2}p-x|\sigma|_{1}}{2}. So we know that

Va+1σ>x2px|σ|12V^{\sigma}_{a+1}>\dfrac{x^{2}p-x|\sigma|_{1}}{2}

Also, VaσV^{\sigma}_{a} and Va+1σV^{\sigma}_{a+1} are relevant coefficients less than or equal to Va+1σV^{\sigma}_{a+1} but not less than or equal to Vaσ1V^{\sigma}_{a}-1. Hence

TVa+1σσTVaσ1σ2T^{\sigma}_{V^{\sigma}_{a+1}}-T^{\sigma}_{V^{\sigma}_{a}-1}\geq 2

Since Va+1σV^{\sigma}_{a+1} is a relevant coefficient, we have TVa+1σσ>TVa+1σ1σT^{\sigma}_{V^{\sigma}_{a+1}}>T^{\sigma}_{V^{\sigma}_{a+1}-1}. By Lemma 6.11, we know that

Va+1σ(Vaσ1)x+1V^{\sigma}_{a+1}-(V^{\sigma}_{a}-1)\geq x+1

Another corollary of Lemma 6.11 lemma is that

Corollary 6.13.

When r2r\geq 2, statement (3) of Lemma 4.2 is an equality. In other words

TV0relrel=TV0relσT^{rel}_{V_{0}^{rel}}=T^{\sigma}_{V_{0}^{rel}}
Proof.

When r=2r=2 and σ=(1)\sigma=(1), statement (2) of Lemma 4.2 says that ν+rel=0\nu^{+rel}=0. Hence V0rel=0V_{0}^{rel}=0 and the result is trivial. Now we assume that it is not this special case.

By Remark 2 after the proof of Lemma 4.2, it suffices to show that

(10) TV0relσTV0rel1σ1T^{\sigma}_{V_{0}^{rel}}-T^{\sigma}_{V_{0}^{rel}-1}\leq 1

We assume that TV0relσ>TV0rel1σT^{\sigma}_{V_{0}^{rel}}>T^{\sigma}_{V_{0}^{rel}-1}, otherwise (10) is trivially true.

Whenever r2r\geq 2, apart from the special case we dealt with at the start of this proof, statement (2) of Lemma 4.2 implies that

ν+rel>p|σ|1\nu^{+rel}>p-|\sigma|_{1}

Lemma 6.7 implies that

V0rel=Vν+relσ>p|σ|12V^{rel}_{0}=V^{\sigma}_{\nu^{+rel}}>\dfrac{p-|\sigma|_{1}}{2}

Now we look at the statement of Lemma 6.11. When a=V0rela=V^{rel}_{0}, b=V0rel1b=V^{rel}_{0}-1, x=1x=1, the first two conditions of Lemma 6.11 are true but the conclusion is false. Therefore, the third condition

TV0relσTV0rel1σ2T^{\sigma}_{V_{0}^{rel}}-T^{\sigma}_{V_{0}^{rel}-1}\geq 2

must be false. ∎

Note that Corollary 6.10 implies that when 0k<p0\leq k<p, Vk+1σ=Vkσ or Vkσ+1V^{\sigma}_{k+1}=V^{\sigma}_{k}\text{ or }V^{\sigma}_{k}+1.

Also, note that Corollary 6.10 and Corollary 6.12 together imply that for any positive integer xx, when 0k<p0\leq k<p, Vxp+k+1σVxp+kσV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k} must be either xx or x+1x+1.

To prove equation (9), by induction on kk, it suffices to show that

Vxp+k+1σVxp+kσ={x+1 if Vk+1σ=Vkσ+1x if Vk+1σ=VkσV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=\begin{cases}x+1\text{ if }V^{\sigma}_{k+1}=V^{\sigma}_{k}+1\\ x\phantom{+11}\text{ if }V^{\sigma}_{k+1}=V^{\sigma}_{k}\end{cases}
Lemma 6.14.

Let xx be a positive integer, 0k<p0\leq k<p. If Vkσ=Vk+1σV^{\sigma}_{k}=V^{\sigma}_{k+1}, then

Vxp+k+1σVxp+kσ=xV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=x
Proof.

By Corollary 6.12, we know that Vxp+k+1σ>Vxp+kσV^{\sigma}_{xp+k+1}>V^{\sigma}_{xp+k}, and therefore TVxp+kσσ=xp+kT^{\sigma}_{V^{\sigma}_{xp+k}}=xp+k.

Let αSVxp+kσ\alpha\in S_{\leq V^{\sigma}_{xp+k}} be such that ασ=TVxp+kσσ=xp+k\alpha\cdot\sigma=T^{\sigma}_{V^{\sigma}_{xp+k}}=xp+k. Since Vxp+kσV^{\sigma}_{xp+k} is a relevant coefficient, we know that TVxp+kσσ>TVxp+kσ1σT^{\sigma}_{V^{\sigma}_{xp+k}}>T^{\sigma}_{V^{\sigma}_{xp+k}-1}, and hence αSVxp+kσ1\alpha\not\in S_{\leq V^{\sigma}_{xp+k}-1}. So, we know that

12jαj(αj+1)=Vxp+kσ\dfrac{1}{2}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)=V^{\sigma}_{xp+k}

Since Vxp+k+1σV^{\sigma}_{xp+k+1} is a relevant coefficient, we know that TVxp+k+1σσ>TVxp+k+1σ1σT^{\sigma}_{V^{\sigma}_{xp+k+1}}>T^{\sigma}_{V^{\sigma}_{xp+k+1}-1}. Since there are no relevant coefficients between Vxp+k+1σV^{\sigma}_{xp+k+1} and Vxp+kσV^{\sigma}_{xp+k}, we also know that TVxp+kσσ=TVxp+k+1σ1σT^{\sigma}_{V^{\sigma}_{xp+k}}=T^{\sigma}_{V^{\sigma}_{xp+k+1}-1}.

Now we try to apply Lemma 6.9 with a=Vxp+k+1σa=V^{\sigma}_{xp+k+1}, b=Vxp+kσb=V^{\sigma}_{xp+k}. We do not have the first condition ax2p+x|σ|12a\leq\tfrac{x^{2}p+x|\sigma|_{1}}{2}. However, in the proof of Lemma 6.9, the only place where the condition is used is to show that there exists some ii such that αi<xσi\alpha_{i}<x\sigma_{i}. If such an ii exists, even if the condition ax2p+x|σ|12a\leq\tfrac{x^{2}p+x|\sigma|_{1}}{2} is not satisfied, the proof of Lemma 6.9 still goes through and we can conclude that

Vxp+k+1σVxp+kσxV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}\leq x

Hence, we can assume that such an ii does not exist. In that case, α=β+xσ\alpha=\beta+x\sigma for some β\beta with all entries being non-negative. β\beta satisfies

βσ\displaystyle\beta\cdot\sigma =(αxσ)σ\displaystyle=(\alpha-x\sigma)\cdot\sigma
=ασxp\displaystyle=\alpha\cdot\sigma-xp
=k\displaystyle=k

Vkσ1V^{\sigma}_{k}-1 is less than the kthk^{th} relevant coefficient VkV_{k}. Therefore, TVkσ1σ<kT^{\sigma}_{V^{\sigma}_{k}-1}<k and we can conclude that

12jβj(βj+1)Vkσ\dfrac{1}{2}\sum\limits_{j}\beta_{j}(\beta_{j}+1)\geq V^{\sigma}_{k}

Now we have

Vxp+kσ\displaystyle V^{\sigma}_{xp+k} =jαj(αj+1)\displaystyle=\sum\limits_{j}\alpha_{j}(\alpha_{j}+1)
=j(βj+xσj)(βj+xσj+1)\displaystyle=\sum\limits_{j}(\beta_{j}+x\sigma_{j})(\beta_{j}+x\sigma_{j}+1)
=x2p+x|σ|12+x(βσ)+12jβj(βj+1)\displaystyle=\dfrac{x^{2}p+x|\sigma|_{1}}{2}+x(\beta\cdot\sigma)+\dfrac{1}{2}\sum\limits_{j}\beta_{j}(\beta_{j}+1)
x2p+x|σ|12+xk+Vkσ\displaystyle\geq\dfrac{x^{2}p+x|\sigma|_{1}}{2}+xk+V^{\sigma}_{k}

Let cc be maximal such that Vkσ=Vk+cσV^{\sigma}_{k}=V^{\sigma}_{k+c}.

By the maximality of cc, TVkσσ=k+cT^{\sigma}_{V^{\sigma}_{k}}=k+c. Let βSVk\beta^{\prime}\in S_{\leq V_{k}} be such that βσ=k+c\beta^{\prime}\cdot\sigma=k+c.

Let α:=β+xσ\alpha^{\prime}:=\beta^{\prime}+x\sigma. Then we have ασ=xp+k+c\alpha^{\prime}\cdot\sigma=xp+k+c. Also

jαj(αj+1)\displaystyle\sum\limits_{j}\alpha^{\prime}_{j}(\alpha^{\prime}_{j}+1) =j(βj+xσj)(βj+xσj+1)\displaystyle=\sum\limits_{j}(\beta^{\prime}_{j}+x\sigma_{j})(\beta^{\prime}_{j}+x\sigma_{j}+1)
=x2p+x|σ|12+x(βσ)+12jβj(βj+1)\displaystyle=\dfrac{x^{2}p+x|\sigma|_{1}}{2}+x(\beta^{\prime}\cdot\sigma)+\dfrac{1}{2}\sum\limits_{j}\beta^{\prime}_{j}(\beta^{\prime}_{j}+1)
x2p+x|σ|12+xk+xc+Vkσ\displaystyle\leq\dfrac{x^{2}p+x|\sigma|_{1}}{2}+xk+xc+V^{\sigma}_{k}
Vxp+kσ+xc\displaystyle\leq V^{\sigma}_{xp+k}+xc

Hence, αSVxp+kσ+xc\alpha^{\prime}\in S_{V^{\sigma}_{xp+k}+xc} and we can conclude that

TVxp+kσ+xcσxp+k+cT^{\sigma}_{V^{\sigma}_{xp+k}+xc}\geq xp+k+c

So, there are at least xp+k+cxp+k+c relevant coefficients less than or equal to Vxp+kσ+xcV^{\sigma}_{xp+k}+xc. Hence, we can conclude that

Vxp+k+cσVxp+kσ+xcV^{\sigma}_{xp+k+c}\leq V^{\sigma}_{xp+k}+xc

which can be arranged into

Vxp+k+cσVxp+kσxcV^{\sigma}_{xp+k+c}-V^{\sigma}_{xp+k}\leq xc

By Corollary 6.12, we know that

Vxp+k+1σVxp+kσ\displaystyle V^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k} x\displaystyle\geq x
Vxp+k+2σVxp+k+1σ\displaystyle V^{\sigma}_{xp+k+2}-V^{\sigma}_{xp+k+1} x\displaystyle\geq x
\displaystyle\dots
Vxp+k+cσVxp+k+c1σ\displaystyle V^{\sigma}_{xp+k+c}-V^{\sigma}_{xp+k+c-1} x\displaystyle\geq x

The only way that Vxp+k+cσVxp+kσxcV^{\sigma}_{xp+k+c}-V^{\sigma}_{xp+k}\leq xc can hold is that all these inequalities are all equalities. ∎

Now we prove equation (9):

Lemma 6.15.

Let xx be a positive integer and 0k<p0\leq k<p. Then

Vxp+kσ=x2p+x|σ|12+xk+VkσV^{\sigma}_{xp+k}=\dfrac{x^{2}p+x|\sigma|_{1}}{2}+xk+V^{\sigma}_{k}
Proof.

First of all, by Corollary 6.6, we know this is true when k=0k=0. Now we consider

Vxp+pσVxpσ\displaystyle V^{\sigma}_{xp+p}-V^{\sigma}_{xp} =(x+1)2p+(x+1)|σ|12x2p+x|σ|12\displaystyle=\dfrac{(x+1)^{2}p+(x+1)|\sigma|_{1}}{2}-\dfrac{x^{2}p+x|\sigma|_{1}}{2}
=xp+p+|σ|12\displaystyle=xp+\dfrac{p+|\sigma|_{1}}{2}

Therefore, by considering Corollary 6.10 and Corollary 6.12, we know that exactly p+|σ|12\tfrac{p+|\sigma|_{1}}{2} values of k{0,,p1}k\in\{0,\dots,p-1\} satisfy Vxp+k+1σVxp+kσ=x+1V^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=x+1, and the rest of the kk values satisfies Vxp+k+1σVxp+kσ=xV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=x.

Since Vpσ=p+|σ|12V^{\sigma}_{p}=\tfrac{p+|\sigma|_{1}}{2}, by Corollary 6.10 we know that exactly p+|σ|12\tfrac{p+|\sigma|_{1}}{2} values of k{0,,p1}k\in\{0,\dots,p-1\} satisfy Vk+1σ=Vkσ+1V^{\sigma}_{k+1}=V^{\sigma}_{k}+1. Therefore, by considering Lemma 6.14, we conclude that for k{0,,p1}k\in\{0,\dots,p-1\},

Vxp+k+1σVxp+kσ=x+1 if and only if Vk+1σ=Vkσ+1V^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=x+1\text{ if and only if }V^{\sigma}_{k+1}=V^{\sigma}_{k}+1

and

Vxp+k+1σVxp+kσ=x if and only if Vk+1σ=VkσV^{\sigma}_{xp+k+1}-V^{\sigma}_{xp+k}=x\text{ if and only if }V^{\sigma}_{k+1}=V^{\sigma}_{k}

Result follows from induction on kk. ∎

Before we prove Theorem 6.1, we make a few more evaluations of relevant coefficients.

Let odd(σ)odd(\sigma) be the number of odd entries in σ\sigma.

Lemma 6.16.

The following five statements are true:
(1) If pp is even, V2p|σ|12σ=p2V^{\sigma}_{\tfrac{2p-|\sigma|_{1}}{2}}=\tfrac{p}{2}.
(2) If pp is odd, V2p|σ|112σ=p12V^{\sigma}_{\tfrac{2p-|\sigma|_{1}-1}{2}}=\tfrac{p-1}{2}.
(3) Vp|σ|12σpodd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}\geq\tfrac{p-odd(\sigma)}{8}
(4) In (3), equality holds if and only if the even entries of σ\sigma can be partitioned into two sets with equal sum.
(5) Vp|σ|12σp+3odd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}\leq\tfrac{p+3odd(\sigma)}{8}

Proof.

Corollary 6.6 and Corollary 6.10 and Corollary 6.12 together implies that for any 0k|σ|10\leq k\leq|\sigma|_{1},

Vpkσ=p+|σ|12kV^{\sigma}_{p-k}=\dfrac{p+|\sigma|_{1}}{2}-k

This implies statements (1) and (2) of Lemma 6.16.

Now we investigate Vp|σ|12σV^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}.

Consider the full rational greedy algorithm that purchases all options strictly less than 12\tfrac{1}{2} coin per point. The amount of coins spent is

σieven(σi21)(σi2)2σiodd(σi12)(σi12+1)2\displaystyle\sum\limits_{\sigma_{i}\ even}\dfrac{\left(\tfrac{\sigma_{i}}{2}-1\right)\left(\tfrac{\sigma_{i}}{2}\right)}{2}-\sum\limits_{\sigma_{i}\ odd}\dfrac{\left(\tfrac{\sigma_{i}-1}{2}\right)\left(\tfrac{\sigma_{i}-1}{2}+1\right)}{2} =σieven(σi28σi4)+σiodd(σi2818)\displaystyle=\sum\limits_{\sigma_{i}\ even}\left(\dfrac{\sigma_{i}^{2}}{8}-\dfrac{\sigma_{i}}{4}\right)+\sum\limits_{\sigma_{i}\ odd}\left(\dfrac{\sigma_{i}^{2}}{8}-\dfrac{1}{8}\right)
=podd(σ)814σievenσi\displaystyle=\dfrac{p-odd(\sigma)}{8}-\dfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}

The amount of points obtained is

σieven(σi21)σi+σiodd(σi12)σi\displaystyle\sum\limits_{\sigma_{i}\ even}\left(\dfrac{\sigma_{i}}{2}-1\right)\sigma_{i}+\sum\limits_{\sigma_{i}\ odd}\left(\dfrac{\sigma_{i}-1}{2}\right)\sigma_{i} =σieven(σi22σi)+σiodd(σi22σi2)\displaystyle=\sum\limits_{\sigma_{i}\ even}\left(\dfrac{\sigma^{2}_{i}}{2}-\sigma_{i}\right)+\sum\limits_{\sigma_{i}\ odd}\left(\dfrac{\sigma^{2}_{i}}{2}-\dfrac{\sigma_{i}}{2}\right)
=p|σ|1212σievenσi\displaystyle=\dfrac{p-|\sigma|_{1}}{2}-\dfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i}

Hence, Tpodd(σ)814σievenσiσ=p|σ|1212σievenσiT^{\sigma}_{\tfrac{p-odd(\sigma)}{8}-\tfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}}=\dfrac{p-|\sigma|_{1}}{2}-\dfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i}.

By also considering the purchase options that are exactly 12\tfrac{1}{2} coin per point, we conclude that for any 0k12σievenσi0\leq k\leq\tfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i},

Tpodd(σ)814σievenσi+kσ,=p|σ|1212σievenσi+k2T^{\sigma,\mathbb{Q}}_{\tfrac{p-odd(\sigma)}{8}-\tfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}+k}=\dfrac{p-|\sigma|_{1}}{2}-\dfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i}+\dfrac{k}{2}

where kk represents the amount of coins spent on purchasing options that exactly 12\tfrac{1}{2} coin per point under a rational greedy algorithm.

Setting k=14σievenσik=\tfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}, we conclude that

Tpodd(σ)8σ,=p|σ|12T^{\sigma,\mathbb{Q}}_{\tfrac{p-odd(\sigma)}{8}}=\dfrac{p-|\sigma|_{1}}{2}

In particular Tpodd(σ)8σ,=Tpodd(σ)8σT^{\sigma,\mathbb{Q}}_{\tfrac{p-odd(\sigma)}{8}}=T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}} if and only if there is an integral implementation of a greedy algorithm that spends exactly 14σievenσi\tfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i} coins on purchasing options that are exactly 12\tfrac{1}{2} coin per point. Such an integral implementation is possible if and only if there is a subset of even σi\sigma_{i}’s that sum to 12σieven\tfrac{1}{2}\sum\limits_{\sigma_{i}\ even}, which is equivalent to saying that the even entries of σ\sigma can be partitioned into two sets with equal sum. Therefore,

Tpodd(σ)8σp|σ|12T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}}\leq\dfrac{p-|\sigma|_{1}}{2}

with equality if and only if the even entries of σ\sigma can be partitioned into two sets with equal sum.

Note that Tpodd(σ)8σ=p|σ|12T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}}=\tfrac{p-|\sigma|_{1}}{2} implies Vp|σ|12σpodd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}\leq\tfrac{p-odd(\sigma)}{8} and Vp|σ|12+1σ>podd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}+1}>\tfrac{p-odd(\sigma)}{8}, which implies Vp|σ|12σ=podd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}=\tfrac{p-odd(\sigma)}{8} because of Corollary 6.10. Also, Tpodd(σ)8σ<p|σ|12T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}}<\tfrac{p-|\sigma|_{1}}{2} implies Vp|σ|12σ>Tpodd(σ)8σV^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}>T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}}. Therefore, this implies statements (3) and (4) of Lemma 6.16.

Since σ\sigma is a changemaker vector, there is some A{0,,n}A\subseteq\{0,\dots,n\} such that iAσi=12σievenσi\sum\limits_{i\in A}\sigma_{i}=\tfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i}. Consider α\alpha defined with

αi={σi21 if σi is even and iAσi2 if σi is even and iAσi12 if σi is odd and iAσi12+1 if σi is odd and iA\alpha_{i}=\begin{cases}\tfrac{\sigma_{i}}{2}-1\text{ if }\sigma_{i}\text{ is even and }i\not\in A\\ \tfrac{\sigma_{i}}{2}\text{ if }\sigma_{i}\text{ is even and }i\in A\\ \tfrac{\sigma_{i}-1}{2}\text{ if }\sigma_{i}\text{ is odd and }i\not\in A\\ \tfrac{\sigma_{i}-1}{2}+1\text{ if }\sigma_{i}\text{ is odd and }i\in A\end{cases}

Then

ασ\displaystyle\alpha\cdot\sigma =p|σ|1212σievenσi+iAσi\displaystyle=\dfrac{p-|\sigma|_{1}}{2}-\dfrac{1}{2}\sum\limits_{\sigma_{i}\ even}\sigma_{i}+\sum\limits_{i\in A}\sigma_{i}
=p|σ|12\displaystyle=\dfrac{p-|\sigma|_{1}}{2}

and

jαj(αj+1)\displaystyle\sum\limits_{j}\alpha_{j}(\alpha_{j}+1) =podd(σ)814σievenσi+σieveniAσi2+σioddiAσi+12\displaystyle=\dfrac{p-odd(\sigma)}{8}-\dfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}+\sum_{\begin{subarray}{c}\sigma_{i}\ even\\ i\in A\end{subarray}}\dfrac{\sigma_{i}}{2}+\sum_{\begin{subarray}{c}\sigma_{i}\ odd\\ i\in A\end{subarray}}\dfrac{\sigma_{i}+1}{2}
=podd(σ)814σievenσi+12iAσi+σioddiA12\displaystyle=\dfrac{p-odd(\sigma)}{8}-\dfrac{1}{4}\sum\limits_{\sigma_{i}\ even}\sigma_{i}+\dfrac{1}{2}\sum_{i\in A}\sigma_{i}+\sum_{\begin{subarray}{c}\sigma_{i}\ odd\\ i\in A\end{subarray}}\dfrac{1}{2}
=podd(σ)8+number of iA such that σi is odd2\displaystyle=\dfrac{p-odd(\sigma)}{8}+\dfrac{\text{number of }i\in A\text{ such that }\sigma_{i}\text{ is odd}}{2}

Let oddAodd_{A} denote the number of iAi\in A such that σi\sigma_{i} is odd. Now we have

Tpodd(σ)8+oddA2σp|σ|12T^{\sigma}_{\tfrac{p-odd(\sigma)}{8}+\tfrac{odd_{A}}{2}}\geq\dfrac{p-|\sigma|_{1}}{2}

Therefore,

Vp|σ|12σp+4oddAodd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}\leq\tfrac{p+4odd_{A}-odd(\sigma)}{8}

Since oddAodd(σ)odd_{A}\leq odd(\sigma), we conclude that

Vp|σ|12σp+3odd(σ)8V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}\leq\tfrac{p+3odd(\sigma)}{8}

Now we can prove Theorem 6.1

Proof or Theorem 6.1.

When r,pr,p are both even, let r=2xr=2x.

ν+rel\displaystyle\nu^{+rel} =2xp|σ|12\displaystyle=\dfrac{2xp-|\sigma|_{1}}{2}
=(x1)p+2p|σ|12\displaystyle=(x-1)p+\dfrac{2p-|\sigma|_{1}}{2}
V0\displaystyle V_{0} =V0rel\displaystyle=V^{rel}_{0}
=V(x1)p+2p|σ|12σ\displaystyle=V^{\sigma}_{(x-1)p+\tfrac{2p-|\sigma|_{1}}{2}}
=(x1)2p+(x1)|σ|12+(x1)2p|σ|12+V2p|σ|12σ\displaystyle=\dfrac{(x-1)^{2}p+(x-1)|\sigma|_{1}}{2}+(x-1)\dfrac{2p-|\sigma|_{1}}{2}+V^{\sigma}_{\tfrac{2p-|\sigma|_{1}}{2}}
=(r21)2p+(r21)|σ|12+(r21)2p|σ|12+p2\displaystyle=\dfrac{\left(\tfrac{r}{2}-1\right)^{2}p+\left(\tfrac{r}{2}-1\right)|\sigma|_{1}}{2}+\left(\dfrac{r}{2}-1\right)\dfrac{2p-|\sigma|_{1}}{2}+\dfrac{p}{2}

which simplifies to V0=r2p8V_{0}=\dfrac{r^{2}p}{8}. Hence, r2p=8V0r^{2}p=8V_{0}.

When rr is even and pp is odd, let r=2xr=2x.

ν+rel\displaystyle\nu^{+rel} =12+2xp|σ|12\displaystyle=-\dfrac{1}{2}+\dfrac{2xp-|\sigma|_{1}}{2}
=(x1)p+2p|σ|112\displaystyle=(x-1)p+\dfrac{2p-|\sigma|_{1}-1}{2}
Vr2\displaystyle V_{\tfrac{r}{2}} =V0rel\displaystyle=V^{rel}_{0}
=V(x1)p+2p|σ|112σ\displaystyle=V^{\sigma}_{(x-1)p+\tfrac{2p-|\sigma|_{1}-1}{2}}
=(x1)2p+(x1)|σ|12+(x1)2p|σ|112+V2p|σ|112σ\displaystyle=\dfrac{(x-1)^{2}p+(x-1)|\sigma|_{1}}{2}+(x-1)\dfrac{2p-|\sigma|_{1}-1}{2}+V^{\sigma}_{\tfrac{2p-|\sigma|_{1}-1}{2}}
=(r21)2p+(r21)|σ|12+(r21)2p|σ|112+p12\displaystyle=\dfrac{\left(\tfrac{r}{2}-1\right)^{2}p+\left(\tfrac{r}{2}-1\right)|\sigma|_{1}}{2}+\left(\dfrac{r}{2}-1\right)\dfrac{2p-|\sigma|_{1}-1}{2}+\dfrac{p-1}{2}

which simplifies to Vr2=r2p8r4V_{\tfrac{r}{2}}=\dfrac{r^{2}p}{8}-\dfrac{r}{4}.

Hence, V0Vr2=r2p8r4V_{0}\geq V_{\tfrac{r}{2}}=\dfrac{r^{2}p}{8}-\dfrac{r}{4} and V0Vr2+r2=r2p8+r4V_{0}\leq V_{\tfrac{r}{2}}+\tfrac{r}{2}=\dfrac{r^{2}p}{8}+\dfrac{r}{4}. This implies

8V02rr2p8V0+2r8V_{0}-2r\leq r^{2}p\leq 8V_{0}+2r

When rr is odd, let r=2x+1r=2x+1.

ν+rel\displaystyle\nu^{+rel} =(2x+1)p|σ|12\displaystyle=\dfrac{(2x+1)p-|\sigma|_{1}}{2}
=xp+p|σ|12\displaystyle=xp+\dfrac{p-|\sigma|_{1}}{2}
V0\displaystyle V_{0} =V0rel\displaystyle=V^{rel}_{0}
=Vxp+p|σ|12σ\displaystyle=V^{\sigma}_{xp+\tfrac{p-|\sigma|_{1}}{2}}
=(x2p+x|σ|12+xp|σ|12+Vp|σ|12σ\displaystyle=\dfrac{(x^{2}p+x|\sigma|_{1}}{2}+x\dfrac{p-|\sigma|_{1}}{2}+V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}
=(r12)2p+(r12)|σ|12+(r12)2p|σ|12+Vp|σ|12σ\displaystyle=\dfrac{\left(\tfrac{r-1}{2}\right)^{2}p+\left(\tfrac{r-1}{2}\right)|\sigma|_{1}}{2}+\left(\dfrac{r-1}{2}\right)\dfrac{2p-|\sigma|_{1}}{2}+V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}

which simplifies to V0=r2pp8+Vp|σ|12σV_{0}=\tfrac{r^{2}p-p}{8}+V^{\sigma}_{\tfrac{p-|\sigma|_{1}}{2}}. Hence,

r2pp8+podd(σ)8V0r2pp8+p+3odd(σ)8\dfrac{r^{2}p-p}{8}+\dfrac{p-odd(\sigma)}{8}\leq V_{0}\leq\dfrac{r^{2}p-p}{8}+\dfrac{p+3odd(\sigma)}{8}

and r2pp8+podd(σ)8=V0\tfrac{r^{2}p-p}{8}+\tfrac{p-odd(\sigma)}{8}=V_{0} if and only if the even entries of σ\sigma can be partitioned into two sets with equal sum. Result follows from rearranging this inequality. ∎

7. Linear changemakers

Greene [9] showed that σ\sigma must take 1 of the 33 forms, including 26 small families which have a simple structure and 7 large families that involves a process called “weight expansion”. In this section, we work on 1 of the 26 small families. We believe that a similar approach can be used on all small families. Unfortunately, we do not see how does this generalizes to the large families.

The small family structure that we investigate in the section has changemakers in the form

σ=(4s+3,2s+1,s+1,s,1,,1s copies of 1’s)\sigma=\begin{pmatrix}4s+3,&2s+1,&s+1,&s,&\smash{\underbrace{\begin{matrix}1,&\dots&,&1\end{matrix}}_{s\text{ copies of }1\text{'s}}}\end{pmatrix}

where s2s\geq 2.

According to Greene’s work [9], this family corresponds to the j3j\leq-3 case in Berge Type X in Rasmussen’s table [16, §6.2].

At the end of this section we prove Theorem 1.11, which states that any knot in S3S^{3} has at most 7 lensbordant surgeries with an associated changemaker coming from this family.

First we use the following lemma to convert that into a statement about the number of possible σ\sigma instead of the number of possible surgeries.

Lemma 7.1.

Given a knot KK in S3S^{3} and a changemaker vector σ\sigma, if σ(1)\sigma\neq(1) or ν+(K)\nu^{+}(K) is not a triangular number, there is at most one lensbordant surgeries on KK associated with σ\sigma. When σ=(1)\sigma=(1) and ν+(K)\nu^{+}(K) is a triangular number, there are at most two lensbordant surgeries on KK associated with σ\sigma.

Remark: When σ=(1)\sigma=(1), being a lensbordant sugery associated with σ\sigma means that the surgery bounds a rational homology ball.

Proof of Lemma 7.1.

Suppose if the r12pr_{1}^{2}p-surgery and the r22pr_{2}^{2}p-surgery on KK are both lensbordant associated with σ\sigma. Let r2r1r_{2}\geq r_{1}.

Theorem 1.8 states that

2ν++2(r1)r2pr|σ|12ν+2\nu^{+}+2(r-1)\geq r^{2}p-r|\sigma|_{1}\geq 2\nu^{+}

Hence,

r12pr1|σ|1r22pr2|σ|12(r21)r_{1}^{2}p-r_{1}|\sigma|_{1}\geq r_{2}^{2}p-r_{2}|\sigma|_{1}-2(r_{2}-1)

which can be rearranged into

0(r2r1)((r2r1)p|σ|1)+(2r1p2)(r2r11)+2r1(p1)0\geq(r_{2}-r_{1})\left((r_{2}-r_{1})p-|\sigma|_{1}\right)+(2r_{1}p-2)(r_{2}-r_{1}-1)+2r_{1}(p-1)

Therefore, either r2=r1r_{2}=r_{1} or (r2=r1+1r_{2}=r_{1}+1 and p=1p=1).

When p=1p=1, we know that

2ν++2(r1)r2r2ν+2\nu^{+}+2(r-1)\geq r^{2}-r\geq 2\nu^{+}

which can be rearranged into

r(r1)2ν+(r1)(r2)2\dfrac{r(r-1)}{2}\geq\nu^{+}\geq\dfrac{(r-1)(r-2)}{2}

Hence, we can determine the value of rr unless ν+\nu^{+} is a triangular number, in which case there can be two values of rr. ∎

Remark: The case with two values of rr can indeed happen. Both the p2p^{2}-surgery and the (p+1)2(p+1)^{2}-surgery on the torus knot Tp,(p+1)T_{p,(p+1)} bound a smooth rational homology ball [2, Th. 1.4].

Given Lemma 7.1, we now focus on the number of σ\sigma’s instead of the number of surgeries.

The majority of the content in this section is for proving the following

Theorem 7.2.

Let s5s\geq 5 and r2r\geq 2. Let KK be a knot in S3S^{3} and there is a (r,p)(r,p)-lensbordant surgery on KK associated with a changemaker vector in the form

σ=(4s+3,2s+1,s+1,s,1,,1s copies of 1’s)\sigma=\begin{pmatrix}4s+3,&2s+1,&s+1,&s,&\smash{\underbrace{\begin{matrix}1,&\dots&,&1\end{matrix}}_{s\text{ copies of }1\text{'s}}}\end{pmatrix}


Let aa be the number of elements in the set {i0296Vi(K)300}\{i\geq 0\mid 296\leq V_{i}(K)\leq 300\}.

Let bb be the biggest value satisfying the property that there are at least a2\dfrac{a}{2} elements in the set {i0Vi(K)=b}\{i\geq 0\mid V_{i}(K)=b\}.

Then

s=b11 or b11+1 or b11+2s=\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor\text{ or }\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor+1\text{ or }\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor+2

From this point onwards, until we finish proving Theorem 7.2, we let

σ=(4s+3,2s+1,s+1,s,1,,1s copies of 1’s)\sigma=\begin{pmatrix}4s+3,&2s+1,&s+1,&s,&\smash{\underbrace{\begin{matrix}1,&\dots&,&1\end{matrix}}_{s\text{ copies of }1\text{'s}}}\end{pmatrix}

and we assume that s5s\geq 5 and r2r\geq 2.

The following lemma converts information about VirelV^{rel}_{i}’s into information about ViV_{i}’s.

Lemma 7.3.

For any m1>m20m_{1}>m_{2}\geq 0,

(Tm1relTm2rel+1)r>|{i0m1Vi>m2}|>(Tm1relTm2rel1)r(T^{rel}_{m_{1}}-T^{rel}_{m_{2}}+1)r>|\{i\geq 0\mid m_{1}\geq V_{i}>m_{2}\}|>(T^{rel}_{m_{1}}-T^{rel}_{m_{2}}-1)r
Proof.

Tm1relTm2relT^{rel}_{m_{1}}-T^{rel}_{m_{2}} is the number of relevant coefficients having values that are both m1\leq m_{1} and >m2>m_{2}. Let TT be maximal such that VTV_{T} is relevant and >m2>m_{2}. Let tt be minimal such that VtV_{t} is relevant and m1\leq m_{1}. Since relevant coefficients having indices rr apart from each other,

Tt=(Tm1relTm2rel1)rT-t=(T^{rel}_{m_{1}}-T^{rel}_{m_{2}}-1)r

Hence, by monotonicity of the VV coefficients, we conclude that

|{i0m1Vi>m2}|>(Tm1relTm2rel1)r|\{i\geq 0\mid m_{1}\geq V_{i}>m_{2}\}|>(T^{rel}_{m_{1}}-T^{rel}_{m_{2}}-1)r

Also, since VT+rm2V_{T+r}\leq m_{2}, and also if trt\geq r then Vtr>m1V_{t-r}>m_{1},

|{i0m1Vi>m2}|\displaystyle|\{i\geq 0\mid m_{1}\geq V_{i}>m_{2}\}| (T+r1)(tr+1)+1\displaystyle\leq(T+r-1)-(t-r+1)+1
=Tt+2r1\displaystyle=T-t+2r-1
=(Tm1relTm2rel+1)r1\displaystyle=(T^{rel}_{m_{1}}-T^{rel}_{m_{2}}+1)r-1

Now we calculate some values of TσT^{\sigma}.

Lemma 7.4.
T295σ=110s+75T^{\sigma}_{295}=110s+75
Proof.

With 295 coins, when s15s\geq 15, an integral rational greedy algorithm up to 5s\tfrac{5}{s} coin per point gives

α=(20,10,5,5,0,)\alpha=(20,10,5,5,0,\dots)

which implies

T295σ=110s+75T^{\sigma}_{295}=110s+75

When 14s514\geq s\geq 5, rational greedy algorithm up to 214s+3\tfrac{21}{4s+3} coin per point gives

α=(20+521,10,5,4,0,)\alpha=(20+\dfrac{5}{21},10,5,4,0,\dots)

which implies

T295σ,=110s+75+15s21<110s+76T^{\sigma,\mathbb{Q}}_{295}=110s+75+\dfrac{15-s}{21}<110s+76

Hence,

T295σ110s+75T^{\sigma}_{295}\leq 110s+75

By considering α=(20,10,5,5,0,)\alpha=(20,10,5,5,0,\dots), we conclude that

T295σ=110s+75T^{\sigma}_{295}=110s+75

Lemma 7.5.
T300σ=110s+80T^{\sigma}_{300}=110s+80
Proof.

By considering α=(20,10,5,5,1,1,1,1,1,0,)\alpha=(20,10,5,5,1,1,1,1,1,0,\dots), we know that

T300σ110s+80T^{\sigma}_{300}\geq 110s+80

Hence, it suffices to show that

T300σ110s+80T^{\sigma}_{300}\leq 110s+80

Let m=12jαj(αj+1)m=\tfrac{1}{2}\sum\limits_{j}\alpha_{j}(\alpha_{j}+1). Let α=(α0,,αn)\alpha=(\alpha_{0},\dots,\alpha_{n}). Then

ασ=α0(4s+3)+α1(2s+1)+α2(s+1)+α3s+(α4++αn)\alpha\cdot\sigma=\alpha_{0}(4s+3)+\alpha_{1}(2s+1)+\alpha_{2}(s+1)+\alpha_{3}s+(\alpha_{4}+\dots+\alpha_{n})

Let

S={αS300ασ=T300σ and α35}S=\{\alpha\in S_{\leq 300}\mid\alpha\cdot\sigma=T^{\sigma}_{300}\text{ and }\alpha_{3}\geq 5\}

First of all, we argue that if the set SS is non-empty, then T300σ=110s+80T^{\sigma}_{300}=110s+80.

If SS is non-empty, pick an α\alpha in SS with minimal α3\alpha_{3}. We know that α25\alpha_{2}\geq 5 because otherwise we can swap α2,α3\alpha_{2},\alpha_{3} to increase ασ\alpha\cdot\sigma without increasing mm. So, α110\alpha_{1}\geq 10 because otherwise we can decrease α2,α3\alpha_{2},\alpha_{3} by 1 and increase α1\alpha_{1} by 1, which does not increase mm and keep ασ\alpha\cdot\sigma unchanged while decreasing α3\alpha_{3}. So, α020\alpha_{0}\geq 20 because otherwise we can decrease α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3} by 1 and increase α0\alpha_{0} by 1 to increase ασ\alpha\cdot\sigma without increasing mm. Notice that

20×212+10×112+5×62+5×62=295\dfrac{20\times 21}{2}+\dfrac{10\times 11}{2}+\dfrac{5\times 6}{2}+\dfrac{5\times 6}{2}=295

Therefore α\alpha must be (20,10,5,5,1,1,1,1,1)(20,10,5,5,1,1,1,1,1) because otherwise mm will be bigger than 300. This gives

T300σ=110s+80T^{\sigma}_{300}=110s+80

Now we consider the case when SS is empty. In this case, all optimal α\alpha’s have α34\alpha_{3}\leq 4. Among these α\alpha’s, pick one with maximal α3\alpha_{3} and satisfies

α4α5\alpha_{4}\geq\alpha_{5}\geq\dots

We know that at most 44 values of α4,α5,\alpha_{4},\alpha_{5},\dots can be non-zero, because otherwise we can decrease (α3+1)(\alpha_{3}+1) of those values by 1 and increase α3\alpha_{3} by 1, which increases α3\alpha_{3} but does not increase mm and does not decrease ασ\alpha\cdot\sigma.

Let σ1=(4,2,1,1,0,)\sigma_{1}=(4,2,1,1,0,\dots) and σ2=(19,9,5,4,1,1,1,1,0,)\sigma_{2}=(19,9,5,4,1,1,1,1,0,\dots). We have

ασ=(s4)ασ1+ασ2\alpha\cdot\sigma=(s-4)\alpha\cdot\sigma_{1}+\alpha\cdot\sigma_{2}

First we investigate σ1\sigma_{1}. It has p1=22p_{1}=22 and |σ1|1=8|\sigma_{1}|_{1}=8. Lemma 6.15 implies

V110σ1=295 and V111σ1=301V^{\sigma_{1}}_{110}=295\text{ and }V^{\sigma_{1}}_{111}=301

Therefore, T300σ1=110T^{\sigma_{1}}_{300}=110.

Now we investigate σ2\sigma_{2}. It has p2=487p_{2}=487 and |σ2|1=41|\sigma_{2}|_{1}=41. Since T2σ2=28T^{\sigma_{2}}_{2}=28 and T3σ2=38T^{\sigma_{2}}_{3}=38, we know that V33σ2=V34σ2=3V^{\sigma_{2}}_{33}=V^{\sigma_{2}}_{34}=3. Lemma 6.15 implies

V520σ2=300 and V521σ1=301V^{\sigma_{2}}_{520}=300\text{ and }V^{\sigma_{1}}_{521}=301

Therefore, T300σ2=520T^{\sigma_{2}}_{300}=520.

Hence,

ασ\displaystyle\alpha\cdot\sigma =(s4)ασ1+ασ2\displaystyle=(s-4)\alpha\cdot\sigma_{1}+\alpha\cdot\sigma_{2}
110(s4)+520\displaystyle\leq 110(s-4)+520
=110s+80\displaystyle=110s+80

Lemma 7.6.
T11s233s+24σ22s222s28T^{\sigma}_{11s^{2}-33s+24}\leq 22s^{2}-22s-28
Proof.

Let αS11s233s+24\alpha\in S_{\leq 11s^{2}-33s+24} be such that ασ=T11s233s+24σ\alpha\cdot\sigma=T^{\sigma}_{11s^{2}-33s+24}.

Let σ1=(4,2,1,1,0,)\sigma_{1}=(4,2,1,1,0,\dots) and σ2=(4s5,2s3,s1,s2,1,,1)\sigma_{2}=(4s-5,2s-3,s-1,s-2,1,\dots,1). We have

ασ=2ασ1+ασ2\alpha\cdot\sigma=2\alpha\cdot\sigma_{1}+\alpha\cdot\sigma_{2}

First we investigate σ1\sigma_{1}. It has p1=22p_{1}=22 and |σ1|1=8|\sigma_{1}|_{1}=8. Since T1σ1=4T^{\sigma_{1}}_{1}=4 and T2σ1=6T^{\sigma_{1}}_{2}=6 and T3σ1=8T^{\sigma_{1}}_{3}=8, we know that V6σ1=2V^{\sigma_{1}}_{6}=2 and V7σ1=3V^{\sigma_{1}}_{7}=3. Lemma 6.15 implies

V22(s2)+6σ1=11s234s+26 and V22(s2)+7σ1=11s233s+25V^{\sigma_{1}}_{22(s-2)+6}=11s^{2}-34s+26\text{ and }V^{\sigma_{1}}_{22(s-2)+7}=11s^{2}-33s+25

Therefore,

T11s233s+24σ1=22(s2)+6=22s38T^{\sigma_{1}}_{11s^{2}-33s+24}=22(s-2)+6=22s-38

Now we investigate σ2\sigma_{2}. Rational greedy algorithm up to 4s64s5\tfrac{4s-6}{4s-5} coin per point gives

α=(4s7+4s74s6,2s4,s2,s3,0,)\alpha=(4s-7+\dfrac{4s-7}{4s-6},2s-4,s-2,s-3,0,\dots)

which implies

T11s233s+24σ2,=22s266s+48+4s74s6T^{\sigma_{2},\mathbb{Q}}_{11s^{2}-33s+24}=22s^{2}-66s+48+\dfrac{4s-7}{4s-6}

which implies

T11s233s+24σ222s266s+48T^{\sigma_{2}}_{11s^{2}-33s+24}\leq 22s^{2}-66s+48

Hence,

ασ\displaystyle\alpha\cdot\sigma =2ασ1+ασ2\displaystyle=2\alpha\cdot\sigma_{1}+\alpha\cdot\sigma_{2}
2(22s38)+22s266s+48\displaystyle\leq 2(22s-38)+22s^{2}-66s+48
=22s222s28\displaystyle=22s^{2}-22s-28

Lemma 7.7.
T11s233s+25σ22s222s24T^{\sigma}_{11s^{2}-33s+25}\geq 22s^{2}-22s-24
Proof.

This follows from observing that when α=(4s6,2s4,s2,s3)\alpha=(4s-6,2s-4,s-2,s-3), we have

αS11s233s+25 and ασ=22s222s24\alpha\in S_{11s^{2}-33s+25}\text{ and }\alpha\cdot\sigma=22s^{2}-22s-24

Now we can prove Theorem 7.2, which states that when s5s\geq 5 and r2r\geq 2, if we let aa be the number of elements in the set {i0296Vi300}\{i\geq 0\mid 296\leq V_{i}\leq 300\} and let bb be the biggest value satisfying the property that there are at least a2\dfrac{a}{2} elements in the set {i0Vi=b}\{i\geq 0\mid V_{i}=b\}, then

s=b11 or b11+1 or b11+2s=\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor\text{ or }\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor+1\text{ or }\left\lfloor\sqrt{\dfrac{b}{11}}\right\rfloor+2
Proof of Theorem 7.2.

First of all, note that p=22s2+31s+11p=22s^{2}+31s+11 and |σ|1=9s+5|\sigma|_{1}=9s+5. So we have

TV0relrel=ν+rel12+12(2p|σ|1)=22s2+532s+8T^{rel}_{V_{0}^{rel}}=\nu^{+rel}\geq-\dfrac{1}{2}+\dfrac{1}{2}(2p-|\sigma|_{1})=22s^{2}+\dfrac{53}{2}s+8

Hence, when r2r\geq 2,

TV0relσ22s2+532s+8T^{\sigma}_{V_{0}^{rel}}\geq 22s^{2}+\dfrac{53}{2}s+8

Notice that when s5s\geq 5, 22s2+532s+822s^{2}+\dfrac{53}{2}s+8 is strictly bigger than 110s+80110s+80 and 22s222s2822s^{2}-22s-28. Lemma 7.5 and Lemma 7.6 implies that 300300 and 11s233s+2411s^{2}-33s+24 are strictly less than V0relV_{0}^{rel}. This implies 295295, 300300, 11s233s+2411s^{2}-33s+24, and 11s233s+2511s^{2}-33s+25 are all less than or equal to V0relV_{0}^{rel}. Hence, in the statements of Lemma 7.4-7.7, TσT^{\sigma} can be replaced by TrelT^{rel}.

Together with Lemma 7.3, we conclude that

6r>a>4r6r>a>4r

and

|{i0Vi=11s233s+25}|>3r|\{i\geq 0\mid V_{i}={11s^{2}-33s+25}\}|>3r

Hence, we know that b11s233s+25b\geq 11s^{2}-33s+25.

To get an upper bound of bb in terms of ss, notice that since

|{i0Vi=b}|a2>2r|\{i\geq 0\mid V_{i}=b\}|\geq\dfrac{a}{2}>2r

Lemma 7.3 implies that TbrelTb1rel2T^{rel}_{b}-T^{rel}_{b-1}\geq 2. Lemma 6.11 implies that

bp|σ|12=11s2+11s+3b\leq\dfrac{p-|\sigma|_{1}}{2}=11s^{2}+11s+3

Result follows from

11(s2)2<11s233s+25b11s2+11s+3<11(s+1)211(s-2)^{2}<11s^{2}-33s+25\leq b\leq 11s^{2}+11s+3<11(s+1)^{2}

We conclude this section by proving Theorem 1.11, which states that any knot in S3S^{3} has at most 7 lensbordant surgeries with an associated changemaker coming from this family.

Proof of Theorem 1.11.

Fix a knot in S3S^{3}.

When r=1r=1, T1rel=4s+3T^{rel}_{1}=4s+3 is the number of ViV_{i}’s with value 1. Therefore, there is at most 1 possible value of ss.

When r2r\geq 2, Theorem 7.2 implies that there are at most 3 possible values of s5s\geq 5.

Hence, together with the possibilities s=2,3,4s=2,3,4, there are at most 7 possible values of ss. Result follows from Lemma 7.1. ∎

8. Knots in the Poincaré homology sphere

In this section, we work on knots in the Poincaré homology sphere (denoted as 𝒫\mathcal{P}) instead of in S3S^{3}. We use Caudell’s work [6] to generalize our work in Section 3 of this paper to knots in 𝒫\mathcal{P}.

We make the following definitions that can be found in [6, §1.4]:

Let aba\leq b be integers that are either both odd or both even. We define PI(a,b)PI(a,b) to be the set of integers between aa and bb (including a,ba,b) that have the same parity (odd or even) as a,ba,b.

For every negative-definite unimodular lattice LL,

m(L):=max{𝔠,𝔠𝔠Char(L)}m(L):=\max\{\langle\mathfrak{c},\mathfrak{c}\rangle\mid\mathfrak{c}\in Char(L)\}
short(L):={𝔠𝔠Char(L) and 𝔠,𝔠=m(L)}short(L):=\{\mathfrak{c}\mid\mathfrak{c}\in Char(L)\text{ and }\langle\mathfrak{c},\mathfrak{c}\rangle=m(L)\}
Short(L):={𝔠𝔠Char(L) and 𝔠,𝔠=m(L)8}Short(L):=\{\mathfrak{c}\mid\mathfrak{c}\in Char(L)\text{ and }\langle\mathfrak{c},\mathfrak{c}\rangle=m(L)-8\}

For every vector vLv\in L,

c(v):=max{𝔠,v𝔠short(L)}c(v):=\max\{\langle\mathfrak{c},v\rangle\mid\mathfrak{c}\in short(L)\}
C(v):=max{𝔠,v𝔠Short(L)}C(v):=\max\{\langle\mathfrak{c},v\rangle\mid\mathfrak{c}\in Short(L)\}
Definition 8.1.

[6, Def 1.6] A vector τ=(s,σ)E8n7\tau=(s,\sigma)\in-E_{8}\oplus-\mathbb{Z}^{n-7} is an E8E_{8}-changemaker if

PI(c(τ),c(τ))={𝔠,τ𝔠short(E8n7)}PI(-c(\tau),c(\tau))=\{\langle\mathfrak{c},\tau\rangle\mid\mathfrak{c}\in short(-E_{8}\oplus-\mathbb{Z}^{n-7})\}

and

PI(c(τ)+2,C(τ)){𝔠,τ𝔠Short(E8n7)}PI(c(\tau)+2,C(\tau))\subset\{\langle\mathfrak{c},\tau\rangle\mid\mathfrak{c}\in Short(-E_{8}\oplus-\mathbb{Z}^{n-7})\}

Caudell noted that short(E8n7)={0}{±1}n7short(-E_{8}\oplus-\mathbb{Z}^{n-7})=\{0\}\oplus\{\pm 1\}^{n-7}, and therefore the first condition of Definition 8.1 is equivalent to saying that σ\sigma satisfies the changemaker condition (as in Definition 3.1, but the entries of σ\sigma are allowed to be 0).

Note that the definition of changemaker in [6] allows the entries to be 0. That is different from Definition 3.1 which comes from [9].

One important result by Caudell [6, Th. 1.19] is that if the canonical negative definite plumbing 4-manifold of L(p,q)L(p,q) embeds in E8n7-E_{8}\oplus-\mathbb{Z}^{n-7} as an orthogonal complement of an E8E_{8}-changemaker, then L(p,q)L(p,q) is a positive integer surgery on one of the knots described by Tange [18] in 𝒫\mathcal{P}. In this section, we use this to generalize the content of Section 3 to L-space knots in 𝒫\mathcal{P}.

The aim of this section is to prove

Theorem 1.12.

Suppose r2pr^{2}p-surgery on a knot K𝒫K\subset\mathcal{P} produces an L-space that is smoothly 2\mathbb{Z}_{2}-homology cobordant to a reduced L(p,q)L(p,q). Let

i(K,r,p)={2g(K)+r if p is odd2g(K) if p is eveni(K,r,p)=\begin{cases}2g(K)+r\text{ if }p\text{ is odd}\\ 2g(K)\quad\ \ \text{ if }p\text{ is even}\end{cases}

If r2pi(K,r,p)r^{2}p\geq i(K,r,p), then L(p,q)L(p,q) must be a positive integer surgery on a knot in 𝒫\mathcal{P}.

If r2p<i(K,r,p)r^{2}p<i(K,r,p), then L(p,q)L(p,q) must be a positive integer surgery on a knot in S3S^{3}.

We use a similar set up as Section 3. In this section, we let KK be an L-space knot in K𝒫K\subset\mathcal{P}. L(p,q)L(p,q) is reduced and is smoothly Z2Z_{2}-homology cobordant to the r2pr^{2}p-surgery on KK which is denoted as Kr2pK_{r^{2}p}. Let WW be the cobordism. Let PP be the canonical negative definite plumbed 4-manifold with boundary L(p,q)L(p,q). Let Wr2pW_{r^{2}p} be the 4-manifold obtained by attaching a r2pr^{2}p-framed 2-handle along K𝒫(𝒫×[0,1])K\subset\mathcal{P}\subset\partial(\mathcal{P}\times[0,1]). Let i(K,r,p)i(K,r,p) be as described in Theorem 1.12. (sorry for using 𝒫\mathcal{P} and PP to denote different things, which came from using notations from both [6] and [1])

Since L(p,q)L(p,q) is smoothly Z2Z_{2}-homology cobordant to Kr2pK_{r^{2}p}, it must be smoothly rational homology cobordant to Kr2pK_{r^{2}p}, and we also know that |H1(W)||H_{1}(W)| is odd.

We also define the surface Σ\Sigma and label the spinc\text{spin}^{\text{c}} structures on Kr2pK_{r^{2}p} in the same way as in Section 3.

Let X=PL(p,q)WX=P\cup_{L(p,q)}W and Z=XKr2pWr2pZ=X\cup_{K_{r^{2}p}}W_{-r^{2}p}.

[Uncaptioned image]

Figure 3. The set up of Section 8.

Since 𝒫\mathcal{P} is an integer homology 3-sphere, Lemma 3.2-3.8 and Definition 3.9 still applies. To use notations similar to [6], we let τ\tau (instead of σ\sigma) to be a generator of the free part of H2(WWr2p)H_{2}(W\cup-W_{r^{2}p}) such that

[Σ]=rτ+torsion[\Sigma]=r\tau+torsion

In Section 3, the first time we used the information that KS3K\subset S^{3} was in the discussion between Lemma 3.8 and Definition 3.9 when we introduced the VV coefficients and how they relate the dd-invariants. So, we start again from there.

Instead of using those ViV_{i} coefficients, we use the torsion coefficients tit_{i} defined as follows. Consider the symmetrized Alexander polynomial of KK

ΔK(T)=jajTj\Delta_{K}(T)=\sum\limits_{j}a_{j}T^{j}

The torsion coefficients are defined as

ti(K):=j1ja|i|+jt_{i}(K):=\sum\limits_{j\geq 1}ja_{|i|+j}

Some properties of torsion coefficients of L-space knots in 𝒫\mathcal{P} addressed in [18] and also mentioned in [5]. For i0i\geq 0, the tit_{i} coefficients form a monotonic decreasing sequence of non-negative integers. Also, for any positive integer kk, whenever |i|k2|i|\leq\tfrac{k}{2},

(11) 22ti(K)=d(Kk,i)d(Uk,i)2-2t_{i}(K)=d(K_{k},i)-d(U_{k},i)

where KkK_{k} is the kk-surgery on K𝒫K\subset\mathcal{P}, and UkU_{k} is the kk-surgery on the unknot in S3S^{3} (which is L(k,k1)L(k,k-1)). The ii behind denotes the spinc\text{spin}^{\text{c}} structure with label ii.

Equation (11) is almost the same as what we saw in Section 3, but with an extra 2 on the left hand side coming from the fact that d(𝒫)=2d(\mathcal{P})=2.

Lemma 8.2.

Let ii be relevant. Then

88ti=max𝔠Char(Z)𝔠,[Σ]+r2p=2i(𝔠2+(n+1))8-8t_{i}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(Z)\\ \langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i\end{subarray}}(\mathfrak{c}^{2}+(n+1))
Proof.

Everything in the proof of Lemma 3.10 still applies, with tit_{i} replacing ViV_{i}, and an extra 8 on the left hand side coming from the extra 2 on the left hand side of equation (11). ∎

Recall that in this section, |H1(W)||H_{1}(W)| is odd. Consider the Mayer-Vietoris sequence

H1(PWr2p)H1(W)H1(Z)H0(L(p,q)Kr2p)injective\rightarrow H_{1}(P\cup-W_{-r^{2}p})\oplus H_{1}(W)\rightarrow H_{1}(Z)\rightarrow H_{0}(L(p,q)\cup K_{r^{2}p})\xrightarrow{injective}

Since H1(PWr2p)=0H_{1}(P\cup-W_{-r^{2}p})=0, we know that |H1(Z)||H_{1}(Z)| divides |H1(W)||H_{1}(W)|. Hence, H1(Z)H_{1}(Z) has no 2-torsion.

Lemma 8.3.

If r2pi(K,r,p)r^{2}p\geq i(K,r,p), the intersection form of ZZ is E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}.

If r2p<i(K,r,p)r^{2}p<i(K,r,p), the intersection form of ZZ is n+1-\mathbb{Z}^{n+1}.

Proof.

According to Scaduto’s work [17, Cor. 1.4], if a smooth compact oriented negative definite 4-manifold with 𝒫\mathcal{P} boundary and b2=n+1b_{2}=n+1 has no 2-torsion in its homology, then its intersection form is either n+1-\mathbb{Z}^{n+1} or E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}.

Since KK is an L-space knot in 𝒫\mathcal{P}, by combining [20, Prop. 3.7], [13, Prop. 3.5], and [16, Prop. 3.1], we know that the genus g(K)g(K) coincides with the leading degree of the symmetrized Alexander polynomial.

By comparing the definition of i(K,r,p)i(K,r,p) and Definition 3.9, and consider the fact that relevant coefficients are monotonic decreasing, we see that r2pi(K,r,p)r^{2}p\geq i(K,r,p) if and only if there exists some relevant ii such that ti=0t_{i}=0.

Hence, it suffices to show that when there exists some relevant ii such that ti=0t_{i}=0, the intersection form of ZZ cannot be n+1-\mathbb{Z}^{n+1}, and show that when the intersection form of ZZ is E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}, there exists some relevant ii such that ti=0t_{i}=0.

Suppose there exists some relevant ii such that ti=0t_{i}=0.

Similar to the argument in [6, §3], observe that for any characteristic vector 𝔠\mathfrak{c} in n+1-\mathbb{Z}^{n+1}, 𝔠2\mathfrak{c}^{2} is at most (n+1)-(n+1). Hence, it is impossible to obtain

8=𝔠2+(n+1)8=\mathfrak{c}^{2}+(n+1)

Therefore, we can conclude that the intersection form of ZZ cannot be n+1-\mathbb{Z}^{n+1}.

Suppose the intersection form of ZZ is E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}.

Let τ=s+σ\tau=s+\sigma, with sE8s\in-E_{8} and σn+1\sigma\in-\mathbb{Z}^{n+1}. By choosing a suitable orthonormal basis, we can assume that all entries of σ\sigma are non-negative.

By considering 𝔠=(0,(1,,1))short(E8n7)\mathfrak{c}=(0,(1,\dots,1))\in short(-E_{8}\oplus-\mathbb{Z}^{n-7}), Lemma 8.2 implies that

t12r(rp|σ|1)=0t_{\tfrac{1}{2}r(rp-|\sigma|_{1})}=0

Note that

p=τ,τ=s,sσ,σp=\langle\tau,\tau\rangle=-\langle s,s\rangle-\langle\sigma,\sigma\rangle

Since E8E_{8} is an even lattice, s,s\langle s,s\rangle is always even. Hence, pp has the same parity (odd or even) as |σ|1|\sigma|_{1}. Therefore, t12r(rp|σ|1)t_{\tfrac{1}{2}r(rp-|\sigma|_{1})} is relevant. ∎

Lemma 8.4.

If the intersection form of ZZ is n+1-\mathbb{Z}^{n+1}, then L(p,q)L(p,q) is a positive integer surgery on a knot in S3S^{3}.

Proof.

This is implied by the argument in Section 3, with tit_{i} replacing ViV_{i}, τ\tau replacing σ\sigma, Lemma 8.2 replacing Lemma 3.10, and in the proof of Lemma 3.11 we consider the relevant coefficients that equals to 1 instead of relevant coefficients that equals to 0. ∎

From this point onwards, until we start proving Theorem 1.12, we assume that the intersection form of ZZ is E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}.

Let τ=s+σ\tau=s+\sigma, with sE8s\in-E_{8} and σn7\sigma\in-\mathbb{Z}^{n-7}. By choosing a suitable orthonormal basis, we can assume that all entries of σ\sigma are non-negative.

We now prove that τ\tau is an E8E_{8}-changemaker. The definition of E8E_{8}-changemaker consists of two conditions. We prove them one by one using arguments similar to the ones in [6, §3].

Lemma 8.5.

σ\sigma is a changemaker (with entries allowed to be 0)

Proof.

We follow a similar argument as in the proof of Lemma 3.11.

When 𝔠=(0,(1,,1))short(E8n7)\mathfrak{c}=(0,(1,\dots,1))\in short(-E_{8}\oplus-\mathbb{Z}^{n-7}), we have 𝔠2+(n+1)=8\mathfrak{c}^{2}+(n+1)=8 and

𝔠,[Σ]+r2p=2(r2pr|σ|12)\langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2\left(\dfrac{r^{2}p-r|\sigma|_{1}}{2}\right)

By Lemma 8.2, we conclude that

t12r(rp|σ|1)=0t_{\tfrac{1}{2}r(rp-|\sigma|_{1})}=0

By Lemma 8.2, we know that for all relevant i12r(rp|σ|1)i\geq\tfrac{1}{2}r(rp-|\sigma|_{1}), there exists some 𝔠Char(E8n7)\mathfrak{c}\in Char(-E_{8}\oplus-\mathbb{Z}^{n-7}) such that 𝔠2=(n7)\mathfrak{c}^{2}=-(n-7) and 𝔠,[Σ]+r2p=2i\langle\mathfrak{c},[\Sigma]\rangle+r^{2}p=2i. Since short(E8n7)={0}{±1}n7short(-E_{8}\oplus-\mathbb{Z}^{n-7})=\{0\}\oplus\{\pm 1\}^{n-7}, we know that such a 𝔠\mathfrak{c} must be in {0}{±1}n7\{0\}\oplus\{\pm 1\}^{n-7}. The rest of the proof of Lemma 3.11 applies. ∎

Lemma 8.6.

If s,s4\langle s,s\rangle\leq-4, then

PI(c(τ)+2,C(τ)){𝔠,τ𝔠Short(E8n7)}PI(c(\tau)+2,C(\tau))\subset\{\langle\mathfrak{c},\tau\rangle\mid\mathfrak{c}\in Short(-E_{8}\oplus-\mathbb{Z}^{n-7})\}
Proof.

From [6, Lemma 3.4], we know that if s,s4\langle s,s\rangle\leq-4, then for all 𝔠Short(E8n7)\mathfrak{c}\in Short(-E_{8}\oplus-\mathbb{Z}^{n-7}), we have |𝔠,τ||τ,τ|=p|\langle\mathfrak{c},\tau\rangle|\leq|\langle\tau,\tau\rangle|=p. Hence, for all 𝔠Short(E8n7)\mathfrak{c}\in Short(-E_{8}\oplus-\mathbb{Z}^{n-7}), we have

(12) 𝔠,τ+rp(r1)p\langle\mathfrak{c},\tau\rangle+rp\geq(r-1)p

Let i0i_{0} be the smallest relevant index such that ti0=0t_{i_{0}}=0. Let i1i_{1} is the smallest relevant index relevant index such that ti11t_{i_{1}}\leq 1.

Case 1: rr is odd or pp is even
Let i:=iri^{\prime}:=\tfrac{i}{r}. Lemma 8.2 can be rewritten as:

For all 0irp20\leq i^{\prime}\leq\tfrac{rp}{2},

88tri=max𝔠Char(Z)𝔠,τ+rp=2i(𝔠2+(n+1))8-8t_{ri^{\prime}}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(Z)\\ \langle\mathfrak{c},\tau\rangle+rp=2i^{\prime}\end{subarray}}(\mathfrak{c}^{2}+(n+1))

Hence, when 0irp20\leq i^{\prime}\leq\tfrac{rp}{2},

tri=0𝔠short such that 𝔠,τ+rp=2it_{ri^{\prime}}=0\Leftrightarrow\exists\mathfrak{c}\in short\text{ such that }\langle\mathfrak{c},\tau\rangle+rp=2i^{\prime}

and

tri=1tri0 and 𝔠Short such that 𝔠,τ+rp=2it_{ri^{\prime}}=1\Leftrightarrow t_{ri^{\prime}}\neq 0\text{ and }\exists\mathfrak{c}\in Short\text{ such that }\langle\mathfrak{c},\tau\rangle+rp=2i^{\prime}

By (12), we know that

i1=r2(rpC(τ)) and i0=r2(rpc(τ))i_{1}=\dfrac{r}{2}\left(rp-C(\tau)\right)\text{ and }i_{0}=\dfrac{r}{2}\left(rp-c(\tau)\right)

Let jPI(c(τ)+2,C(τ))j\in PI(c(\tau)+2,C(\tau)). We want to show that j{𝔠,τ𝔠Short(E8n7)}j\in\{\langle\mathfrak{c},\tau\rangle\mid\mathfrak{c}\in Short(-E_{8}\oplus-\mathbb{Z}^{n-7})\}. Let

i=r2(rpj)i=\dfrac{r}{2}\left(rp-j\right)

Since jj has the same parity (odd or even) as C(τ)C(\tau), ii differs from i1i_{1} by a multiple of rr. So, ii is relevant. Since i1i<i0i_{1}\leq i<i_{0}, we must have ti=1t_{i}=1. Hence, there exists some 𝔠Short\mathfrak{c}\in Short such that

𝔠,τ+rp=2ir\langle\mathfrak{c},\tau\rangle+rp=\dfrac{2i}{r}

This implies 𝔠,τ=j\langle-\mathfrak{c},\tau\rangle=j.

Case 2: rr is even and pp is odd
Let i:=ir12i^{\prime}:=\tfrac{i}{r}-\tfrac{1}{2}. Lemma 8.2 can be rewritten as:

For all 0irp210\leq i^{\prime}\leq\tfrac{rp}{2}-1,

88tri=max𝔠Char(Z)𝔠,τ+rp=2i+1(𝔠2+(n+1))8-8t_{ri^{\prime}}=\max_{\begin{subarray}{c}\mathfrak{c}\in Char(Z)\\ \langle\mathfrak{c},\tau\rangle+rp=2i^{\prime}+1\end{subarray}}(\mathfrak{c}^{2}+(n+1))

Since rr is even, (12) implies that for all 𝔠Short\mathfrak{c}\in Short, we have

𝔠,τ+rp10\langle\mathfrak{c},\tau\rangle+rp-1\geq 0

The rest of the proof of Case 1 applies. ∎

Lemma 8.7.

τ\tau is an E8E_{8}-changemaker.

Proof.

According to the arguments in the proofs of [6, Prop. 3.6] and [6, Prop. 4.2], τ\tau is an E8E_{8}-changemaker when s,s=0,2\langle s,s\rangle=0,-2. Since E8E_{8} is an even lattice, the only other possibility is s,s4\langle s,s\rangle\leq-4, which Lemma 8.6 says that τ\tau is an E8E_{8}-changemaker. Hence, in all cases, τ\tau is an E8E_{8}-changemaker. ∎

Now we can prove Theorem 1.12.

Proof of Theorem 1.12.

By Lemma 3.3, Lemma 8.7, and [6, Th. 1.19], we conclude that when the intersection form of ZZ is E8n7-E_{8}\oplus-\mathbb{Z}^{n-7}, L(p,q)L(p,q) is a positive integer surgery on a knot in the Poincaré homology sphere described by Tange. The rest of the proof of Theorem 1.12 follows from Lemma 8.3 and Lemma 8.4. ∎

References

  • [1] Paolo Aceto, Daniele Celoria, and JungHwan Park, Rational cobordisms and integral homology, Compositio Mathematica 156 (2020), 1825 – 1845.
  • [2] Paolo Aceto and Marco Golla, Dehn surgeries and rational homology balls, Algebraic & Geometric Topology 17 (2017), no. 1, 487 – 527.
  • [3] Paolo Aceto, Marco Golla, Kyle Larson, and Ana G. Lecuona, Surgeries on torus knots, rational balls, and cabling, 2020.
  • [4] John Berge, Some knots with surgeries yielding lens spaces, unpublished manuscript.
  • [5] Jacob Caudell, Three lens space summands from the Poincaré homology sphere, 2021.
  • [6] by same author, Dehn surgery on knots in the Poincaré homology sphere, Ph.D. thesis, Boston College, 2023.
  • [7] Marc Culler, Cameron Gordon, John Luecke, and Peter Shalen, Dehn surgery on knots, Annals of Mathematics 125 (1985), 237 – 300.
  • [8] Simon K. Donaldson, The orientation of Yang-Mills moduli spaces and 4-manifold topology, Journal of Differential Geometry 26 (1987), no. 3, 397 – 428.
  • [9] Joshua E. Greene, The lens space realization problem, Annals of Mathematics 177 (2010), 449–511.
  • [10] by same author, L-space surgeries, genus bounds, and the cabling conjecture, Journal of Differential Geometry 100 (2015), no. 3, 491 – 506.
  • [11] Jennifer Hom and Zhongtao Wu, Four-ball genus bounds and a refinement of the Ozsváth-Szabó tau-invariant, Journal of Symplectic Geometry 14 (2016), no. 1, 305 – 323.
  • [12] Duncan McCoy, Bounds on alternating surgery slopes, Algebraic & Geometric Topology 17 (2017), no. 5, 2603–2634.
  • [13] Yi Ni, A note on knot Floer homology of links, Geometry & Topology 10 (2006), no. 2, 695 – 713.
  • [14] Yi Ni and Zhongtao Wu, Cosmetic surgeries on knots in s3s^{3}, J. Reine Angew. Math. 706 (2015), 1–17.
  • [15] Jacob Rasmussen, Floer homology and knot complements, Ph.D. thesis, Harvard University, 2003.
  • [16] by same author, Lens space surgeries and L-space homology spheres, 2007.
  • [17] Christopher Scaduto, On definite lattices bounded by a homology 3-sphere and Yang-Mills instanton Floer theory, 2018.
  • [18] Motoo Tange, Lens spaces given from L-space homology 3-spheres, Experimental Mathematics 18 (2009), 285 – 301.
  • [19] by same author, Homology spheres yielding lens spaces, Proceedings of the Gökova Geometry-Topology Conference (2017), 73–121.
  • [20] Zhongtao Wu, Cosmetic surgery in integral homology L-spaces, Geometry & Topology 15 (2011), no. 2, 1157 – 1168.