11.07.2015 Views

Knot invariants and their applications to constructions of quasi ...

Knot invariants and their applications to constructions of quasi ...

Knot invariants and their applications to constructions of quasi ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Knot</strong> <strong>invariants</strong> <strong>and</strong> <strong>their</strong> <strong>applications</strong><strong>to</strong> <strong>constructions</strong> <strong>of</strong> <strong>quasi</strong>-morphismson groupsMichael Br<strong>and</strong>enbursky


<strong>Knot</strong> <strong>invariants</strong> <strong>and</strong> <strong>their</strong> <strong>applications</strong><strong>to</strong> <strong>constructions</strong> <strong>of</strong> <strong>quasi</strong>-morphismson groupsResearch ThesisIn partial fulfillment <strong>of</strong> the requirements <strong>of</strong> theDegree <strong>of</strong> Doc<strong>to</strong>r <strong>of</strong> PhilosophyMichael Br<strong>and</strong>enburskySubmitted <strong>to</strong> the Senate <strong>of</strong> theTechnion - Israel Institute <strong>of</strong> TechnologyIyar 5770 Haifa May 2010


The research was carried out under the supervision <strong>of</strong> Pr<strong>of</strong>essorMichael En<strong>to</strong>v <strong>and</strong> Pr<strong>of</strong>essor Michael Polyak in the Departmen<strong>to</strong>f Mathematics.I wish <strong>to</strong> thank my advisors for suggesting the main problems,<strong>and</strong> for <strong>their</strong> constant help, both mathematical <strong>and</strong> nonmathematical,<strong>and</strong> encouragement during last four years.I also owe thanks <strong>to</strong> other people:To Pr<strong>of</strong>essor Leonid Polterovich for many very helpful comments,<strong>and</strong> for giving me the opportunity <strong>to</strong> present a part <strong>of</strong> this workat the University <strong>of</strong> Chicago.To Pr<strong>of</strong>essor Michah Sageev for suggesting the pro<strong>of</strong> <strong>of</strong> Lemma1.3.5.To Sergei Lanzat for many valuable discussions.To my parents <strong>and</strong> my sister for <strong>their</strong> constant support.I wish <strong>to</strong> thank my wonderful wife Carolina, whom I love verymuch, for her endless support <strong>and</strong> for being there for me for allthese years.The generous financial support <strong>of</strong> the Technion—Israel Institute<strong>of</strong> Technology is gratefully acknowledged.


ContentsAbstract 1Introduction. 31 Preliminaries in knot theory <strong>and</strong> symplectic geometry 71.1 Classical <strong>invariants</strong> <strong>of</strong> knots <strong>and</strong> links . . . . . . . . . . . . . . . 71.1.1 <strong>Knot</strong> signatures . . . . . . . . . . . . . . . . . . . . . . . 71.1.2 Concordance group, four ball genus <strong>and</strong> genus <strong>of</strong> a knot . 91.1.3 Rasmussen invariant s . . . . . . . . . . . . . . . . . . . 91.1.4 Ozsvath-Szabo invariant τ . . . . . . . . . . . . . . . . . 121.2 Polynomial <strong>and</strong> Vassiliev <strong>invariants</strong><strong>of</strong> classical <strong>and</strong> virtual links . . . . . . . . . . . . . . . . . . . . 131.2.1 Conway <strong>and</strong> HOMFLYPT polynomials . . . . . . . . . . 131.2.2 Long classical <strong>and</strong> virtual links . . . . . . . . . . . . . . 141.2.3 Finite type <strong>invariants</strong> <strong>of</strong> classical <strong>and</strong> virtual links . . . . 161.3 Braid groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191.4 Quasi-morphisms . . . . . . . . . . . . . . . . . . . . . . . . . . 201.5 Calabi homomorphism . . . . . . . . . . . . . . . . . . . . . . . 212 Quasi-morphisms on braid groups defined by knot <strong>invariants</strong> 232.1 Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 232.2 Examples <strong>and</strong> <strong>applications</strong> . . . . . . . . . . . . . . . . . . . . . 272.2.1 Rasmussen <strong>quasi</strong>-morphism . . . . . . . . . . . . . . . . 272.2.2 Connection between τ, s <strong>and</strong> braid number . . . . . . . . 302.2.3 Signature <strong>quasi</strong>-morphisms . . . . . . . . . . . . . . . . . 313 Generalized Gambaudo-Ghys construction 363.1 Gambaudo-Ghys construction . . . . . . . . . . . . . . . . . . . 36


3.2 Reeb graphs <strong>and</strong> computations forau<strong>to</strong>nomous diffeomorphisms . . . . . . . . . . . . . . . . . . . . 403.2.1 Angle-action coordinates <strong>and</strong> the Reeb graph . . . . . . 413.2.2 Main Theorem . . . . . . . . . . . . . . . . . . . . . . . 433.3 Properties <strong>of</strong> induced <strong>quasi</strong>-morphisms on the group D . . . . . 493.4 Induced <strong>quasi</strong>-morphisms <strong>and</strong> Calabihomomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . 543.4.1 Rasmussen <strong>quasi</strong>-morphisms on D . . . . . . . . . . . . . 543.4.2 Signature <strong>quasi</strong>-morphisms <strong>and</strong> Calabihomomorphism . . . . . . . . . . . . . . . . . . . . . . . 553.5 Open problems . . . . . . . . . . . . . . . . . . . . . . . . . . . 584 Link <strong>invariants</strong> via counting surfaces 594.1 Gauss diagrams <strong>and</strong> arrow diagrams . . . . . . . . . . . . . . . 594.1.1 Gauss diagrams <strong>of</strong> classical <strong>and</strong> virtual links . . . . . . . 604.1.2 Arrow diagrams <strong>and</strong> Gauss diagram formulas . . . . . . 624.1.3 Surfaces corresponding <strong>to</strong> arrow diagrams . . . . . . . . 634.1.4 Ascending <strong>and</strong> descending arrow diagrams . . . . . . . . 644.1.5 Separating states . . . . . . . . . . . . . . . . . . . . . . 664.2 Counting surfaces with one boundarycomponent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 684.2.1 Invariants <strong>of</strong> long links . . . . . . . . . . . . . . . . . . . 694.2.2 Properties <strong>of</strong> A n,m (G) <strong>and</strong> D n,m (G) . . . . . . . . . . . . 714.2.3 Alex<strong>and</strong>er-Conway polynomials <strong>of</strong> long virtual links . . . 744.3 Counting surfaces with two boundarycomponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.3.1 Link <strong>invariants</strong> <strong>and</strong> diagrams with two boundary components. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.3.2 Properties <strong>of</strong> I n . . . . . . . . . . . . . . . . . . . . . . . 864.3.3 The case <strong>of</strong> knots . . . . . . . . . . . . . . . . . . . . . . 90Bibliography 92


List <strong>of</strong> Figures1.1 Smoothing <strong>of</strong> a crossing . . . . . . . . . . . . . . . . . . . . . . 71.2 Seifert surface for the Hopf link obtained by the Seifert algorithm 81.3 Closure ̂α <strong>of</strong> a braid α . . . . . . . . . . . . . . . . . . . . . . . 121.4 Conway triple . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131.5 Reidemeister moves . . . . . . . . . . . . . . . . . . . . . . . . . 151.6 Virtual Reidemeister moves . . . . . . . . . . . . . . . . . . . . 151.7 Connected sum <strong>of</strong> long trefoil knot <strong>and</strong> long Hopf link . . . . . . 161.8 Diagrams <strong>of</strong> classical, long, virtual <strong>and</strong> long virtual trefoil knots 161.9 Genera<strong>to</strong>r σ i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192.1 The braid α = σ 1 · . . . · σ n−1 . . . . . . . . . . . . . . . . . . . . 282.2 Braid η i,n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323.1 Reeb graph which corresponds <strong>to</strong> the function H with 3 maximumpoints <strong>and</strong> 2 saddle points . . . . . . . . . . . . . . . . . . 433.2 The braid η i,k,n . . . . . . . . . . . . . . . . . . . . . . . . . . . 453.3 Curve c <strong>to</strong>gether with annuli A l <strong>and</strong> A l+1 . . . . . . . . . . . . . 474.1 Diagrams <strong>of</strong> based <strong>and</strong> long classical Hopf links <strong>to</strong>gether withthe associated Gauss diagrams . . . . . . . . . . . . . . . . . . . 604.2 Virtual trefoil <strong>and</strong> a virtual Hopf link with the correspondingGauss diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . 614.3 Based <strong>and</strong> long virtual trefoil with the corresponding Gauss diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614.4 Reidemeister moves <strong>of</strong> Gauss diagrams . . . . . . . . . . . . . . 624.5 Connected arrow diagrams . . . . . . . . . . . . . . . . . . . . . 624.6 Constructing a surface from an arrow diagram . . . . . . . . . . 644.7 Ascending <strong>and</strong> descending labeling . . . . . . . . . . . . . . . . 674.8 Smoothing <strong>of</strong> an arrow . . . . . . . . . . . . . . . . . . . . . . . 67


4.9 A Conway triple <strong>of</strong> Gauss diagrams . . . . . . . . . . . . . . . . 714.10 Dependence on a basepoint . . . . . . . . . . . . . . . . . . . . . 724.11 The virtualization move . . . . . . . . . . . . . . . . . . . . . . 754.12 Diagrams <strong>of</strong> long virtual knots that differ by an application <strong>of</strong>the virtualization move . . . . . . . . . . . . . . . . . . . . . . . 764.13 A diagram <strong>of</strong> the Kishino knot . . . . . . . . . . . . . . . . . . . 764.14 Gauss diagrams <strong>of</strong> the Kishino knot . . . . . . . . . . . . . . . . 774.15 A mirror pair <strong>of</strong> Gauss diagrams . . . . . . . . . . . . . . . . . . 784.16 A virtual knot <strong>and</strong> two <strong>of</strong> its based Gauss diagrams . . . . . . . 794.17 Identifying ascending separating states containing α l . . . . . . 824.18 Identifying other ascending separating states containing α l . . . 824.19 Comparison <strong>of</strong> ascending separating states <strong>of</strong> G <strong>and</strong> ˜G . . . . . 834.20 Comparison <strong>of</strong> other ascending separating states <strong>of</strong> G <strong>and</strong> ˜G . . 844.21 Gauss diagrams G, G 1 <strong>and</strong> G 2 . . . . . . . . . . . . . . . . . . . 854.22 Correspondence <strong>of</strong> separating states <strong>of</strong> G 0 <strong>and</strong> G ± . . . . . . . . 884.23 Link H 2 <strong>and</strong> its Gauss diagram . . . . . . . . . . . . . . . . . . 904.24 Two versions <strong>of</strong> the Reidemeister move Ω 1 . . . . . . . . . . . . 91


AbstractIn this work we study knot <strong>and</strong> link <strong>invariants</strong> <strong>and</strong> <strong>their</strong> <strong>applications</strong> <strong>to</strong> <strong>constructions</strong><strong>of</strong> <strong>quasi</strong>-morphisms on braids groups <strong>and</strong> the group D <strong>of</strong> area-preservingcompactly supported diffeomorphisms <strong>of</strong> a two-dimensional open disc.In the first part <strong>of</strong> the thesis, which is independent from the subsequentchapters, we give a sufficient criterion for a knot/link invariant <strong>to</strong> define <strong>quasi</strong>morphismson braid groups, provided a certain way <strong>of</strong> closing braids in<strong>to</strong> knotsor links. We then discuss a generalized Gambaudo-Ghys construction whichallows <strong>to</strong> build <strong>quasi</strong>-morphisms on D. In particular, we study <strong>quasi</strong>-morphismson braid groups <strong>and</strong> D defined in this way by knot <strong>and</strong> link <strong>invariants</strong> comingfrom the knot Floer homology <strong>and</strong> Khovanov-type link homology. We alsocompute the values <strong>of</strong> <strong>quasi</strong>-morphisms obtained by this construction on thetime-one flow <strong>of</strong> a generic time-independent Hamil<strong>to</strong>nian H in terms <strong>of</strong> theReeb graph <strong>of</strong> H.In the second part <strong>of</strong> the thesis we consider link <strong>invariants</strong> arising fromthe Conway <strong>and</strong> HOMFLYPT polynomials. It is known that any Vassilievinvariant may be obtained from a Gauss diagram formula that involves counting(with signs <strong>and</strong> multiplicities) subdiagrams <strong>of</strong> certain geometric-combina<strong>to</strong>rialtypes. These formulas generalize the calculation <strong>of</strong> a linking number by countingsigns <strong>of</strong> crossings in a link diagram. Until recently, explicit formulas <strong>of</strong> thistype were known only for few <strong>invariants</strong> <strong>of</strong> low degrees. We generalize theresult <strong>of</strong> Chmu<strong>to</strong>v-CapKhoury-Rossi <strong>and</strong> present Gauss diagram formulas for allcoefficients <strong>of</strong> ∇(L), where L is a link with arbitrary number <strong>of</strong> components, <strong>and</strong>∇(L) is the Conway polynomial <strong>of</strong> L. We discuss an interesting interpretation<strong>of</strong> these formulas in terms <strong>of</strong> counting surfaces <strong>of</strong> a certain genus <strong>and</strong> with oneboundary component. We also present two different extensions <strong>of</strong> the Conwaypolynomial <strong>to</strong> long virtual links. We compare these extensions with the existingextensions <strong>of</strong> the Alex<strong>and</strong>er <strong>and</strong> Conway polynomials <strong>and</strong> show that they arenew.1


In the remaining part <strong>of</strong> the thesis we modify the Chmu<strong>to</strong>v-CapKhoury-Rossi construction <strong>and</strong> present Gauss diagram formulas for the coefficients <strong>of</strong>the first partial derivative <strong>of</strong> the HOMFLYPT polynomial, with respect <strong>to</strong> thevariable a, evaluated at a = 1. These formulas are related, in a similar way, <strong>to</strong>a certain count <strong>of</strong> orientable surfaces with two boundary components. At theend we present a modification <strong>of</strong> these formulas in case <strong>of</strong> knots.2


Introduction.<strong>Knot</strong> theory is a field <strong>of</strong> mathematics which studies knots <strong>and</strong> links in 3-manifolds. Being <strong>of</strong> its own interest, knot theory plays an important role inthe study <strong>of</strong> braid groups <strong>and</strong> low-dimensional <strong>to</strong>pology <strong>and</strong> has many <strong>applications</strong>in symplectic geometry <strong>and</strong> other mathematical fields. Invariants <strong>of</strong> linksin the 3-sphere, S 3 , are basic <strong>to</strong>ols <strong>of</strong> knot theory which have been extensivelystudied for many years.Among the most intriguing link <strong>invariants</strong> there are Alex<strong>and</strong>er, Jones <strong>and</strong>HOMFLYPT polynomials. The Alex<strong>and</strong>er polynomial is a classical invarian<strong>to</strong>f links in S 3 (see e.g. [2], [40]). For many years, before the discovery <strong>of</strong> theJones polynomial in the 1980s (see e.g. [33], [34]), it was thought <strong>to</strong> be theunique polynomial invariant, i.e a polynomial whose coefficients are link <strong>invariants</strong>.The HOMFLYPT polynomial (see e.g. [24], [41], [54]), discoveredshortly after the Jones polynomial, is a 2-variable polynomial which reduces<strong>to</strong> Alex<strong>and</strong>er <strong>and</strong> Jones 1-variable polynomials by certain changes <strong>of</strong> variables.Recently new homological knot <strong>invariants</strong> were discovered by Ozsvath-Szabo[49], Rasmussen [55], Khovanov [36] <strong>and</strong> Khovanov-Rozansky [37]. These homologicaltheories generalize Alex<strong>and</strong>er, Jones <strong>and</strong> HOMFLYPT polynomialsin the following sense: <strong>their</strong> graded Euler characteristics are equal (up <strong>to</strong> somenormalization) <strong>to</strong> Alex<strong>and</strong>er, Jones <strong>and</strong> HOMFLYPT polynomials respectively.In the first part <strong>of</strong> the thesis we discuss link <strong>invariants</strong>, arising from knotFloer homology <strong>and</strong> Khovanov homology. We investigate some <strong>of</strong> the <strong>applications</strong><strong>of</strong> these <strong>invariants</strong> in the study <strong>of</strong> braid groups <strong>and</strong> in symplectic geometry.In the second part <strong>of</strong> the thesis we establish a connection between <strong>invariants</strong>coming from the HOMFLYPT polynomial <strong>and</strong> orientable surfaces.The results in the first part are based on a connection between link <strong>invariants</strong><strong>and</strong> <strong>quasi</strong>-morphisms on braid groups. Recall that a <strong>quasi</strong>-morphism on a groupG is a function φ : G → R such that for some K φ ≥ 0 one has|φ(ab) − φ(a) − φ(b)| ≤ K φ3


for all a, b ∈ G. A <strong>quasi</strong>-morphism φ is called homogeneous if φ(a m ) = mφ(a)for all a ∈ G <strong>and</strong> m ∈ Z. Any <strong>quasi</strong>-morphism φ can be homogenized: setting˜φ(a) :=limk→+∞ φ(ak )/k (1)we get a homogeneous (possibly trivial) <strong>quasi</strong>-morphism ˜φ. Quasi-morphismsare known <strong>to</strong> be a helpful <strong>to</strong>ol in the study <strong>of</strong> non-Abelian groups, especially theones that admit a few or no (linearly independent) real-valued homomorphisms.In this work we discuss <strong>quasi</strong>-morphisms on the groups P n <strong>of</strong> pure braids onn strings. These <strong>quasi</strong>-morphisms on P n are constructed from certain <strong>invariants</strong><strong>of</strong> n-component links in R 3 in the following way: close up a pure braid <strong>to</strong> alink in the st<strong>and</strong>ard way <strong>and</strong> take the value <strong>of</strong> the invariant on that link. Weformulate sufficient conditions for a knot invariant <strong>to</strong> yield a <strong>quasi</strong>-morphismon the full braid group B n on n strings <strong>and</strong> hence on P n . We also considertwo specific remarkable knot/link <strong>invariants</strong>: the Rasmussen link invariant s[56], which comes from a Khovanov-type theory, <strong>and</strong> the Ozsvath-Szabo knotinvariant τ [50], which comes from the knot Floer homology. In [4] Baader hasshown that the Rasmussen link invariant s defines a <strong>quasi</strong>-morphism on B n . Weshow that the homogenization <strong>of</strong> this <strong>quasi</strong>-morphism is equal <strong>to</strong> the classicallinking number homomorphism on B n . We also show that the situation withthe Ozsvath-Szabo invariant τ is similar: it defines a <strong>quasi</strong>-morphism on B n<strong>and</strong> its homogenization is again the linking number homomorphism multipliedby 2. We also present an inequality which connects s, τ <strong>and</strong> the braid number<strong>of</strong> a knot.Later we discuss the group D <strong>of</strong> compactly supported area-preserving diffeomorphisms<strong>of</strong> a unit open two-dimensional disc in the Euclidean plane. Thegroup D admits a unique (continuous, in the proper sense) homomorphism<strong>to</strong> the reals – the famous Calabi homomorphism (see e.g. [5], [13], [25]). Atthe same time D is known <strong>to</strong> admit many (linearly independent) homogeneous<strong>quasi</strong>-morphisms (see e.g. [8], [26], [11]). In this thesis we consider a particularconstruction <strong>of</strong> such <strong>quasi</strong>-morphisms, due <strong>to</strong> Gambaudo <strong>and</strong> Ghys [26]. This isan explicit geometric construction which produces <strong>quasi</strong>-morphisms on D from<strong>quasi</strong>-morphisms on the groups P n . To show that the homogenizations <strong>of</strong> theresulting <strong>quasi</strong>-morphism on D is non-trivial, we follow Gambaudo <strong>and</strong> Ghys<strong>and</strong> evaluate it on the diffeomorphism generated by a certain time-independentHamil<strong>to</strong>nian function on the disc.We discuss the <strong>applications</strong> <strong>of</strong> the Gambaudo-Ghys construction <strong>to</strong> the<strong>quasi</strong>-morphisms on P n defined by τ, s <strong>and</strong> ω-signature knot/link <strong>invariants</strong>.4


In the first two cases the resulting homogeneous <strong>quasi</strong>-morphisms on D are homomorphisms,equal, up <strong>to</strong> a non-zero multiplicative constant, <strong>to</strong> the Calabihomomorphism. In the third case many <strong>of</strong> the resulting homogeneous <strong>quasi</strong>morphismsare not homomorphisms.We also discuss the computation <strong>of</strong> the <strong>quasi</strong>-morphisms on D, obtainedby Gambaudo-Ghys construction, on diffeomorphisms generated by time-independent(compactly supported) Hamil<strong>to</strong>nians. For a generic Hamil<strong>to</strong>nian H<strong>of</strong> this sort we present the result <strong>of</strong> the computation in terms <strong>of</strong> the Reebgraph <strong>of</strong> H <strong>and</strong> the integral <strong>of</strong> the push-forward <strong>of</strong> H <strong>to</strong> the graph against acertain signed measure on the graph. This result enables us <strong>to</strong> show that theCalabi homomorphism <strong>and</strong> the <strong>quasi</strong>-morphism on D induced by the signatureinvariant <strong>of</strong> n-component links are asymp<strong>to</strong>tically equivalent, as n → ∞, onthe flows generated by time-independent (compactly supported) Hamil<strong>to</strong>nians.In the second part <strong>of</strong> the thesis, based on a joint work with M. Polyak, weconsider link <strong>invariants</strong> arising from the Conway <strong>and</strong> HOMFLYPT polynomials.The HOMFLYPT polynomial P (L) is an invariant <strong>of</strong> an oriented link L (seee.g. [24], [41], [54]). It is a Laurent polynomial in two variables a <strong>and</strong> z, whichsatisfies the following skein relation:( ) ( ) ( )aP − a −1 P = zP .+The HOMFLYPT polynomial is normalized(in)the following way. If O m is anm−1.m-component unlink, then P (O m ) =The Conway polynomial−a−a −1z∇ may be defined as ∇(L) := P (L)| a=1 . This polynomial is a renormalizedversion <strong>of</strong> the Alex<strong>and</strong>er polynomial (see e.g. [20], [40]). All coefficients <strong>of</strong> ∇are so-called finite type or Vassiliev <strong>invariants</strong>.One <strong>of</strong> the mainstream <strong>and</strong> simplest techniques for producing Vassiliev <strong>invariants</strong>are so-called Gauss diagram formulas (see [30], [53]). These formulasgeneralize the calculation <strong>of</strong> a linking number by counting subdiagrams <strong>of</strong>special geometric-combina<strong>to</strong>rial types with signs <strong>and</strong> weights in a given linkdiagram. This technique is also very helpful in the rapidly developing field <strong>of</strong>virtual knot theory (see [35]), as well as in 3-manifold theory (see [44]).Until recently, explicit formulas <strong>of</strong> this type were known only for few <strong>invariants</strong><strong>of</strong> low degrees. The situation has changed with works <strong>of</strong> Chmu<strong>to</strong>v-CapKhoury-Rossi [15] <strong>and</strong> Chmu<strong>to</strong>v-Polyak [17]. In [15] Chmu<strong>to</strong>v-CapKhoury-Rossi presented an infinite family <strong>of</strong> Gauss diagram formulas for all coefficients<strong>of</strong> ∇(L), where L is a knot or a two-component oriented link. We explain how50


each formula for the coefficient c n <strong>of</strong> z n is related <strong>to</strong> a certain count <strong>of</strong> orientablesurfaces <strong>of</strong> a certain genus, <strong>and</strong> with one boundary component. The genus dependsonly on n <strong>and</strong> the number <strong>of</strong> the components <strong>of</strong> L. These formulas maybe viewed as a certain combina<strong>to</strong>rial analog <strong>of</strong> Gromov-Witten <strong>invariants</strong>.In this work we generalize the result <strong>of</strong> Chmu<strong>to</strong>v-CapKhoury-Rossi <strong>to</strong> linkswith arbitrary number <strong>of</strong> components. We present a direct pro<strong>of</strong> <strong>of</strong> this result,without any prior assumption on the existence <strong>of</strong> the Conway polynomial. Itenables us <strong>to</strong> present two different extensions <strong>of</strong> the Conway polynomial <strong>to</strong>long virtual links. We compare these extensions with the existing versions <strong>of</strong>the Alex<strong>and</strong>er <strong>and</strong> Conway polynomials for virtual links, <strong>and</strong> show that theyare new. In particular, we give a new pro<strong>of</strong> <strong>of</strong> the fact that the famous Kishinoknot K T [38] is non-classical, by calculating these polynomials for K T .This leads <strong>to</strong> a natural question: how <strong>to</strong> produce link <strong>invariants</strong> by countingorientable surfaces with an arbitrary number <strong>of</strong> boundary components? In thisthesis we deal with a model case, when the number <strong>of</strong> boundary componentsis two. We modify Chmu<strong>to</strong>v-CapKhoury-Rossi construction <strong>and</strong> present aninfinite family <strong>of</strong> Gauss diagram formulas for the coefficients <strong>of</strong> the first partialderivative <strong>of</strong> the HOMFLYPT polynomial, w.r.t. the variable a, evaluated ata = 1. This family is related, in a similar way, <strong>to</strong> the family <strong>of</strong> orientablesurfaces with two boundary components. We conjecture that a similar coun<strong>to</strong>f orientable surfaces with n boundary components (up <strong>to</strong> some normalization)will give an infinite family <strong>of</strong> Gauss diagram formulas for the coefficients <strong>of</strong> then−1 derivative, w.r.t. the variable a in the HOMFLYPT polynomial, evaluatedat a = 1. At the end we present a modification <strong>of</strong> these formulas in case <strong>of</strong>knots.A challenging problem is <strong>to</strong> try <strong>to</strong> build a theory similar <strong>to</strong> knot Floer homology,which will categorify the Conway polynomial. While we get the Conwaypolynomial from the Alex<strong>and</strong>er polynomial by the change z = t − 1 2 − t 1 2 <strong>of</strong> variables,it is not clear how <strong>to</strong> ”change variables” in knot Floer homology. It seemsplausible that the first step in this direction would be a better underst<strong>and</strong>ing <strong>of</strong>the connection between orientable surfaces <strong>and</strong> the coefficients <strong>of</strong> the Conwaypolynomial. We hope that the techniques described in this thesis will help <strong>to</strong>deal with this problem.6


Chapter 1Preliminaries in knot theory <strong>and</strong>symplectic geometryIn this chapter we shall recall the basic notions <strong>and</strong> facts in knot theory <strong>and</strong>symplectic geometry that will be repeatedly used in the following chapters.1.1 Classical <strong>invariants</strong> <strong>of</strong> knots <strong>and</strong> linksIn this section we recall some classical <strong>and</strong> new <strong>invariants</strong> <strong>of</strong> knots <strong>and</strong> links inS 3 .1.1.1 <strong>Knot</strong> signaturesIn this subsection we recall the definition <strong>of</strong> the signature link invariant.Let L be an oriented link in S 3 , then there exists an oriented surface S Lwith boundary L. Each such orientable surface S L is called a Seifert surface <strong>of</strong>L. For the reader’s convenience we recall the Seifert algorithm for construction<strong>of</strong> Seifert surfaces [40]. Let D be a diagram <strong>of</strong> an oriented link L. Let usFigure 1.1: Smoothing <strong>of</strong> a crossing7


smoothen each crossing in D as shown in Figure 1.1. The resulting smootheddiagram ̂D consists <strong>of</strong> oriented simple closed curves, which are called the Seifertcircles. Thus ̂D is the boundary <strong>of</strong> a union <strong>of</strong> disjoint discs. We join these discs<strong>to</strong>gether with half-twisted strips corresponding <strong>to</strong> the crossings in the diagram.This yields an oriented surface bounded by L. If this surface is disconnected,then we connect its components by the connected sum operation.Example 1.1.1. In Figure 1.2 we present a diagram <strong>of</strong> the positive Hopf linkL <strong>to</strong>gether with a Seifert surface S L obtained by the Seifert algorithm.Figure 1.2: Seifert surface for the Hopf link obtained by the Seifert algorithmWe choose a basis {a 1 , . . . , a 2g+|L|−1 } in H 1 (S L , Z). There exists a symmetricbilinear form on H 1 (S L , Z), which is defined as follows:Ω(a i , a j ) := lk(a i , a + j ) + lk(a j, a + i ),where lk is the linking number <strong>and</strong> a + i is a push <strong>of</strong> the curve, which representsa i in S L , from S L along the positive normal direction <strong>to</strong> S L . Tensoring by Rwe get a symmetric bilinear form on H 1 (S L , R). The signature <strong>of</strong> this form isindependent <strong>of</strong> the choices <strong>of</strong> the Seifert surface S L <strong>and</strong> the basis <strong>of</strong> H 1 (S L , Z),see [40]. Thus it is an invariant <strong>of</strong> L <strong>and</strong> is denoted by sign(L).For any complex number ω ≠ 1 <strong>and</strong> link L there exists ω-signature linkinvariant sign ω (L), such that sign −1 (L) = sign(L). It is defined as follows. Wetensor the bilinear form Ω by C, <strong>and</strong> we get a bilinear form on H 1 (S L , C). Thesignature <strong>of</strong> the following hermitian form on H 1 (S L , C)Ω ω (a i , a j ) = (1 − ω)Ω(a i , a j ) + (1 − ω)Ω(a j , a i )is independent <strong>of</strong> the choice <strong>of</strong> the Seifert surface S L <strong>and</strong> the base for H 1 (S L , Z),see [40]. Thus sign ω (L) is a link invariant for each ω ≠ 1.8


1.1.2 Concordance group, four ball genus <strong>and</strong> genus <strong>of</strong>a knotLet K be a knot in S 3 . The genus <strong>of</strong> a knot K is the minimal genus <strong>of</strong> anySeifert surface S K . It is denoted by g(K). The four-ball genus <strong>of</strong> a knot K in S 3is the minimal genus <strong>of</strong> an oriented surface with boundary, which is embeddedin D 4 such that its boundary is a knot K ⊂ S 3 = ∂D 4 . It is denoted by g 4 (K).Note that g 4 (K) ≤ g(K).For any knot K, the knot K ∗ denotes its mirror image, <strong>and</strong> the knot −Kdenotes the same knot with the reversed orientation. For any two knots K 0<strong>and</strong> K 1 the knot K 0 #K 1 represents <strong>their</strong> connected sum. We say that K 0 isequivalent <strong>to</strong> K 1 , if there exists an embedding Ψ : S 1 × [0, 1] → S 3 × [0, 1]such that Ψ(S 1 × {0}) = K 0 <strong>and</strong> Ψ(S 1 × {1}) = K 1 . We denote by Conc(S 3 )the Abelian group whose elements are equivalent classes <strong>of</strong> knots in S 3 , <strong>and</strong>the multiplication is the connected sum operation. It is easy <strong>to</strong> see that themultiplication is well-defined. The following lemma is a well-known fact in knottheory, see e.g. [40, 57]. For the reader’s convenience we will prove this lemma.Lemma 1.1.2. For any knot K the knot K# − K ∗ bounds an embedded discin D 4 .Pro<strong>of</strong>. Let Ψ : S 1 × [0, 1] → S 3 × [0, 1] be an embedding such thatΨ(S 1 × {t}) = K × {t}for every t ∈ [0, 1]. There exists an arc ξ in S 1 such that Ψ((S 1 \ ξ) × [0, 1]) ishomeomorphic <strong>to</strong> a disc D 2 <strong>and</strong> bounds K# − K ∗ in D 4 . We can smooth thecorners <strong>of</strong> (S 1 \ξ)×[0, 1] so that Im(Ψ) <strong>of</strong> the resulting manifold is diffeomorphic<strong>to</strong> D 2 <strong>and</strong> bounds K# − K ∗ in D 4 .It follows from Lemma 1.1.2 that for any knot K the equivalence class <strong>of</strong> theknot −K ∗ represents the inverse element in Conc(S 3 ). The identity element inConc(S 3 ) is the equivalence class <strong>of</strong> the unknot.A knot K in S 3 = ∂D 4 is called slice if there exists an embedding <strong>of</strong> the unitdisc D 2 ⊂ R 2 in D 4 such that the image <strong>of</strong> an embedding <strong>of</strong> ∂D 2 is the knot K.1.1.3 Rasmussen invariant sThe Rasmussen invariant s(K) <strong>of</strong> a knot K in S 3 was discovered by JacobRasmussen in 2004, see [56]. It comes from the Lee theory [39] which is closely9


elated <strong>to</strong> the Khovanov homology [36]. This is a very powerful knot invariantwhich, for example, gives a lower bound on the four-ball genus <strong>of</strong> a knot. Itwas also used by Rasmussen <strong>to</strong> give the first combina<strong>to</strong>rial pro<strong>of</strong> <strong>of</strong> the Milnorconjecture, which states that g(K) = g 4 (K) for any <strong>to</strong>rus knot K. Recently thisinvariant was extended <strong>to</strong> links by Beliakova <strong>and</strong> Wehrli [10]. For the reader’sconvenience we will recall the definition <strong>and</strong> some properties <strong>of</strong> s.Let D be a diagram <strong>of</strong> an oriented link L. In [36] Khovanov associated <strong>to</strong>D a bigraded chain complex ( ⊕ C i,j (D), d) over Q, such thati,jd : C i,j (D) → C i+1,j (D).It turns out that the homology groups <strong>of</strong> this complex are <strong>invariants</strong> <strong>of</strong> L <strong>and</strong>the graded Euler characteristic is equal <strong>to</strong> the unnormalized Jones polynomial.For more information about Khovanov homology see e.g. [6, 62].In [39] Lee associated <strong>to</strong> D a chain complex CLee ∗ (D) := (⊕ C i (D), d Lee )iover Q, such that d Lee : C i (D) → C i+1 (D) <strong>and</strong> C i (D) = ⊕ C i,j (D). Shejproved that the homology <strong>of</strong> this complex is an invariant <strong>of</strong> L. In addition sheshowed that Rank(HLee ∗ (L)) = 2m , where m is the number <strong>of</strong> components <strong>of</strong> L.She also showed that any orientation θ <strong>of</strong> L defines a cycle s θ in CLee ∗ (D) <strong>and</strong>there is an explicit isomorphismH ∗ Lee(L) ∼ = Q{[s θ ]| θ is an orientation <strong>of</strong> L}.In Lee theory the differential d Lee does not respect the j grading. Everyelement u ∈ C i (D) can be written as a sum <strong>of</strong> monomials u = u 1 + · · · + u k .We setj(u) := min{j(u i )| i = 1, . . . , k}.Thus j defines a filtration on CLee ∗ (D) by settingF r C ∗ Lee(D) := {u ∈ C ∗ Lee(D)| j(u) ≥ r}.It follows from the definition <strong>of</strong> d Lee that it is a filtered map for a filtration F ron H ∗ Lee (L) defined as follows: a class x ∈ H∗ Lee (L) is in F r , if <strong>and</strong> only if, it hasa representative which is an element <strong>of</strong> F r C ∗ Lee (D). Thus for each x ∈ H∗ Lee (L)we defines(x) := max{j(u)| [u] = x}<strong>and</strong> F r := {x ∈ H ∗ Lee(L)| s(x) ≥ r}.10


For an oriented knot K defines min (K) := min{s(x)| x ∈ H ∗ Lee(K), x ≠ 0},s max (K) := max{s(x)| x ∈ H ∗ Lee(K), x ≠ 0}.The Rasmussen invariant s(K) is defined as followss(K) := s min(K) + s max (K).2It turns out that s(K) = s min (K) + 1.If θ is an orientation <strong>of</strong> L then we denote by θ the reversed orientation <strong>of</strong>each component <strong>of</strong> L. In [10] Beliakova <strong>and</strong> Wehrli defineds(L) := min(s([s θ ] + [s θ]), s([s θ ] − [s θ])) + 1.It follows from Lemma 3.5 in [56] that the above invariant coincides with theRasmussen invariant s when L is a knot.It has many interesting properties. We list here eight <strong>of</strong> them:s(K) = s(−K), (1.1)s(K 1 #K 2 ) = s(K 1 ) + s(K 2 ), (1.2)s(K ∗ ) = −s(K), (1.3)|s(K)| ≤ 2g 4 (K), (1.4)s(K − ) ≤ s(K + ) ≤ s(K − ) + 2, (1.5)where K + <strong>and</strong> K − are knots that differ by a single crossing change: from apositive crossing in K + <strong>to</strong> a negative one in K − (positive/negative crossings areshown in Figure 1.4a).For an alternating knot K we haveFor a positive link Ls(K) = sign(K). (1.6)s(L) = w(D(L)) − o(D(L)) + 1, (1.7)11


where D(L) is some positive diagram <strong>of</strong> L, w(D(L)) is the number <strong>of</strong> positivecrossings in D(L) minus the number <strong>of</strong> negative crossings in D(L), <strong>and</strong> o(D(L))is the number <strong>of</strong> Seifert circles. When the Seifert algorithm is applied <strong>to</strong> D(K)we have the following inequality1 + w(D(K)) − o(D(K)) ≤ s(K) ≤ −1 + w(D(K)) + o(D(K)), (1.8)where D(K) is any diagram <strong>of</strong> a knot K, see [4]. In fact, it is shown in [4] thatproperty (1.8) follows from properties (1.3), (1.5) <strong>and</strong> (1.7).01 01 01α01 01 01Figure 1.3: Closure ̂α <strong>of</strong> a braid α1.1.4 Ozsvath-Szabo invariant τIn [50] Ozsvath <strong>and</strong> Szabo defined a knot invariant τ which comes from the knotFloer homology [43, 49, 55]. This invariant satisfies in particular the followingproperties:τ(K ∗ ) = −τ(K), (1.9)τ(K 1 #K 2 ) = τ(K 1 ) + τ(K 2 ), (1.10)τ(−K) = τ(K), (1.11)|τ(K)| ≤ g 4 (K), (1.12)0 ≤ τ(K + ) − τ(K − ) ≤ 1, (1.13)τ(̂α) = w(̂α) − n + 1 , (1.14)212


where α ∈ B n is a positive braid (see Definition 1.3.2), such that ̂α is a knot.By ̂α we denote the link in S 3 which is obtained in the natural way from α, seeFigure 1.3. The first five properties were proved in the paper <strong>of</strong> Ozsvath <strong>and</strong>Szabo [50], <strong>and</strong> the last property was proved by Livings<strong>to</strong>n in [42].1.2 Polynomial <strong>and</strong> Vassiliev <strong>invariants</strong><strong>of</strong> classical <strong>and</strong> virtual linksIn this section we recall the definitions <strong>of</strong> some knot polynomials <strong>and</strong> finite type<strong>invariants</strong> <strong>of</strong> classical <strong>and</strong> virtual knots.1.2.1 Conway <strong>and</strong> HOMFLYPT polynomialsConway polynomialThe Conway polynomial ∇(L) is an invariant <strong>of</strong> an oriented link L, see e.g.[24, 41, 54]. It is a polynomial in variable z, which satisfies the following skeinrelation. Denote by L + , L − <strong>and</strong> L 0 a triple <strong>of</strong> links with diagrams which are+ −a0bFigure 1.4: Conway tripleidentical except for a small fragment, where L + <strong>and</strong> L − have a positive <strong>and</strong> anegative crossing respectively, <strong>and</strong> L 0 has a smoothed crossing, see Figure 1.4.Such a triple <strong>of</strong> links is called a Conway triple. The Conway polynomial ∇(L)satisfies∇(L + ) − ∇(L − ) = z∇(L 0 )It is normalized by ∇(O) = 1, where O is the unknot. A changez = t − 1 2 − t 1 2 <strong>of</strong> variables turns ∇(L) in<strong>to</strong> the Alex<strong>and</strong>er polynomial ∆(L). Formore information about these polynomials see [20, 40].13


HOMFLYPT polynomialThe HOMFLYPT polynomial P (L) is an invariant <strong>of</strong> an oriented link L, seee.g. [24, 41, 54]. It is a Laurent polynomial in two variables a <strong>and</strong> z whichsatisfies the following skein relation:aP (L + ) − a −1 P (L − ) = zP (L 0 ). (1.15)The HOMFLYPT is normalized by P (O m ) =( )a−a m−1, −1z where Om is anm-component unlink.Remark 1.2.1. Note that P (L)| a=1 = ∇(L) for any oriented link L in S 3 .1.2.2 Long classical <strong>and</strong> virtual linksDefinition 1.2.2. By a long m-component link diagram we mean a smoothimmersion f : R ⊔ S 1 ⊔ . . . ⊔ S 1 → R 2 such that1) it coincides with a st<strong>and</strong>ard embedding <strong>of</strong> y-axis outside <strong>of</strong> the unit discD 2 ⊂ R 2 ;2) each intersection point is double <strong>and</strong> transverse;3) each intersection point is endowed with classical (with a choice for underpass<strong>and</strong> overpass specified) crossing structure.Definition 1.2.3. A long link is an equivalence class <strong>of</strong> long link diagramsmodulo Reidemeister moves, see Figure 1.5.Definition 1.2.4. A virtual link diagram with m components is a generic immersion<strong>of</strong> m disjoint circles in<strong>to</strong> the plane, with double points divided in<strong>to</strong>real crossing points <strong>and</strong> virtual crossing points, with the real crossing pointsenhanced by information on overpasses <strong>and</strong> underpasses (as for classical linkdiagrams). At a virtual crossing the branches are not divided in<strong>to</strong> an overpass<strong>and</strong> an underpass.A virtual crossing is denoted by. Two virtual link diagrams are equivalentif they may be transformed one <strong>to</strong> another by finite sequence <strong>of</strong> Reidemeistermoves <strong>and</strong> virtual Reidemeister moves, see Figure 1.5 <strong>and</strong> Figure 1.6respectively. Two virtual link diagrams represent the same virtual link, if onecan be obtained from the other by a sequence <strong>of</strong> these moves.14


Figure 1.5: Reidemeister movesFigure 1.6: Virtual Reidemeister movesDefinition 1.2.5. By a long virtual m-component link diagram we mean asmooth immersion f : R ⊔ S 1 ⊔ . . . ⊔ S 1 → R 2 such that1) it coincides with a st<strong>and</strong>ard embedding <strong>of</strong> y-axis outside <strong>of</strong> the unit discD 2 ⊂ R 2 ;2) each intersection point is double <strong>and</strong> transverse;3) each intersection point is endowed with classical (with a choice for underpass<strong>and</strong> overpass specified) or virtual crossing structure.Definition 1.2.6. A long virtual link is an equivalence class <strong>of</strong> long virtual linkdiagrams modulo Reidemeister <strong>and</strong> virtual Reidemeister moves.Definition 1.2.7. Let L 1 <strong>and</strong> L 2 two long (virtual) links, then the connectedsum L 1 #L 2 is a concatenation <strong>of</strong> L 1 <strong>and</strong> L 2 : just place a diagram <strong>of</strong> L 2 aftera diagram <strong>of</strong> L 1 , see Figure 1.7.Example 1.2.8. Figure 1.8 shows diagrams <strong>of</strong> classical, long, virtual <strong>and</strong> longvirtual trefoil knots.15


# =Figure 1.7: Connected sum <strong>of</strong> long trefoil knot <strong>and</strong> long Hopf linkFigure 1.8: Diagrams <strong>of</strong> classical, long, virtual <strong>and</strong> long virtual trefoil knotsOur motivation <strong>to</strong> study virtual links comes from the study <strong>of</strong> links in thickenedsurfaces <strong>of</strong> higher genus. Let Σ g be a surface <strong>of</strong> genus g <strong>and</strong> [0, 1] be theunit interval. The knot theory in Σ g × [0, 1] is represented by diagrams drawnon Σ g taken up <strong>to</strong> the usual Reidemeister moves transferred <strong>to</strong> diagrams on thissurface. Abstract <strong>invariants</strong> <strong>of</strong> virtual links can be interpreted as <strong>invariants</strong> <strong>of</strong>links that are embedded in Σ g × [0, 1] for some genus g, see e.g. [35].1.2.3 Finite type <strong>invariants</strong> <strong>of</strong> classical <strong>and</strong> virtual linksIn this subsection we recall definitions <strong>of</strong> finite type or Vassiliev <strong>invariants</strong> <strong>of</strong>classical <strong>and</strong> virtual knots. All these definitions are naturally extended <strong>to</strong> links.16


We start with a case <strong>of</strong> classical knots.Finite type <strong>invariants</strong> <strong>of</strong> classical knotsDefinition 1.2.9. Consider immersed closed curves in R 3 with n transversedouble points, i.e. each such immersion is an embedding when restricted <strong>to</strong>the complement <strong>of</strong> the preimages <strong>of</strong> these double points. Equivalence classes <strong>of</strong>such curves under the action <strong>of</strong> iso<strong>to</strong>pies <strong>of</strong> R 3 are called singular knots with ntransverse double points.Let G be an Abelian group. A G-valued knot invariant v is extended <strong>to</strong>singular knots by induction on the number <strong>of</strong> double points using the inductionformula:( ) ( ) ( )v = v − v . (1.16)Definition 1.2.10. A G-valued knot invariant v is said <strong>to</strong> be <strong>of</strong> finite type orVassiliev, if for some n ∈ N it vanishes for any knot K with more than n doublepoints. A minimal such n is called the degree <strong>of</strong> v.Examples 1.2.11. (i) Let c n be the coefficient <strong>of</strong> z n in the Conway polynomial∇. It follows from [7] that c n is a finite type invariant <strong>of</strong> degree n.(ii) Another family <strong>of</strong> finite type <strong>invariants</strong> is derived from the HOMFLYPTpolynomial in the following way. We make a substitution a = e h <strong>and</strong> take theTaylor expansion in h. The result will be a Laurent polynomial in z <strong>and</strong> apower series in h. Let p k,n be its coefficient at h k z n . Goussarov proved in [29]that p k,n is a finite type invariant <strong>of</strong> degree ≤ k + n.The extension <strong>of</strong> Vassiliev <strong>invariants</strong> <strong>to</strong> virtual knots is not unique. In thefollowing two subsections we recall the definitions <strong>of</strong> two different extensions <strong>of</strong>finite type <strong>invariants</strong> <strong>to</strong> virtual knots.Kauffman FTIIn [35] Kauffman defined the class <strong>of</strong> 4-valent graphs modulo rigid vertex iso<strong>to</strong>py(see [35] for the precise definition). A G-valued virtual knot invariant v isextended <strong>to</strong> this class by induction on the number <strong>of</strong> vertices using the inductionformula (1.16).Definition 1.2.12. A G-valued virtual knot invariant v is said <strong>to</strong> be <strong>of</strong> graphicalfinite type or Vassiliev invariant, if for some n ∈ N it vanishes for any 4-valentgraph Γ with more than n vertices. A minimal such n is called the degree <strong>of</strong> v.+17−


Note that this definition says nothing about the number <strong>of</strong> virtual crossingsin any such Γ.Example 1.2.13. The coefficient at x n in the power series expansion <strong>of</strong> thesubstitution a = e x <strong>of</strong> the Jones-Kauffman polynomial is known <strong>to</strong> be <strong>of</strong> Kauffmanfinite-type <strong>of</strong> degree ≤ n, see [35].Goussarov-Polyak-Viro FTIIn [30] Goussarov, Polyak <strong>and</strong> Viro introduced another extension <strong>of</strong> Vassiliev<strong>invariants</strong> <strong>to</strong> virtual links. It can be described as follows. They introduced anew kind <strong>of</strong> crossing, which is called semi-virtual. At a semi-virtual crossingthere are still over- <strong>and</strong> under-passes. In a diagram a semi-virtual crossingis shown as a real one, but surrounded by a small circle. Virtual knots areextended <strong>to</strong> knots having semi-virtual crossings. Any virtual knot invariant vcan be extended <strong>to</strong> this larger class by applying the following formal relation:( ) ( ) ( )v = v − v . (1.17)Definition 1.2.14. A virtual knot invariant v is called <strong>of</strong> GPV finite-type, iffor some n ∈ N it vanishes for any virtual knot K with more than n semi-virtualcrossings. The minimal such n is called the degree <strong>of</strong> v.Example 1.2.15. Let L be a two component virtual link. The invariantlk 1/2 (L) (lk 2/1 (L) respectively) is defined <strong>to</strong> be the sum <strong>of</strong> the signs <strong>of</strong> thereal crossings <strong>of</strong> L where the first component (respectively the second component)passes over the second one (respectively the first one). Both lk 1/2 <strong>and</strong>lk 2/1 are <strong>of</strong> GPV finite-type <strong>of</strong> degree one.Remark 1.2.16. Equations (1.16) <strong>and</strong> (1.17) imply:( ) ( ) ( )v = v − v . (1.18)It follows that for any GPV-finite type invariant, its restriction <strong>to</strong> classical knotsis a finite type invariant (<strong>of</strong> at most the same degree) in the classical sense.Note that every GPV finite-type invariant <strong>of</strong> degree ≤ n is <strong>of</strong> Kauffmanfinite-type <strong>of</strong> degree ≤ n, but the converse in general is not true, see [35].18


1.3 Braid groupsIn this section we recall a definition <strong>and</strong> some properties <strong>of</strong> braid groups.Definition 1.3.1. The full braid group B n on n strings is abstractly definedvia the following presentation:B n = ⟨σ 1 , . . . , σ n−1 | Rel i,j , if |i − j| ≥ 2, <strong>and</strong> Rel i for 1 ≤ i ≤ n − 2⟩,where Rel i,j <strong>and</strong> Rel i are, respectively, the relations σ i σ j = σ j σ i <strong>and</strong> σ i σ i+1 σ i =σ i+1 σ i σ i+1 .Definition 1.3.2. A braid α ∈ B n is called positive if it can be written as aproduct <strong>of</strong> only non-negative powers <strong>of</strong> the st<strong>and</strong>ard genera<strong>to</strong>rs <strong>of</strong> B n .Each genera<strong>to</strong>r σ i <strong>of</strong> B n has a geometric interpretation, see Figure 1.9.Definition 1.3.3. The braid length l(γ) <strong>of</strong> γ ∈ B n is the length <strong>of</strong> the shortestword representing γ with respect <strong>to</strong> the genera<strong>to</strong>rs σ 1 , . . . , σ n−1 .ii+1Figure 1.9: Genera<strong>to</strong>r σ iThe full braid group B n can be identified with the fundamental groupπ 1 (Y n , z) <strong>of</strong> the space Y n <strong>of</strong> the unordered n-tuples <strong>of</strong> distinct points in D 2 ,where z = (z 1 , . . . , z n ) ∈ Y n is a base point in Y n . Each closed path α(t) ∈ Y n ,such that α(0) = α(1) = z, is given by n-tuple <strong>of</strong> pathes (α 1 (t), . . . , α n (t)) in(D 2 ) ×n , such that α i (t) ≠ α j (t)) for all i ≠ j, α i (0) = z i <strong>and</strong> for each i thereexists unique j such that α i (1) = z j . Thus each braid [α(t)] ∈ π 1 (Y n , z) definesan element s α in a permutation group S n by setting s α (z i ) = α i (1). Thisassignment defines a surjective homomorphism B n → S n . The kernel <strong>of</strong> thishomomorphism is called the pure braid group <strong>and</strong> is denoted by P n . Note thatP n can be identified with the fundamental group π 1 (X n , z) <strong>of</strong> the space X n <strong>of</strong>the ordered n-tuples <strong>of</strong> distinct points in D 2 , where z = (z 1 , . . . , z n ) ∈ X n is abase point in X n .19


Proposition 1.3.4 ([18]). The center Z(B n ) <strong>of</strong> B n coincides with the centerZ(P n ) <strong>of</strong> P n . It is a cyclic group generated by the full twist (σ 1 · . . . · σ n−1 ) n .The following lemma will be needed in Section 3.1.Lemma 1.3.5. Let S be any finite generating set for P n . Then there exist tworeal positive constants K 1,S <strong>and</strong> K 2,S which depend only on S, n <strong>and</strong> σ 1 , . . . , σ n−1so thatl S (γ) ≤ K 1,S · l(γ) + K 2,S ,where l S (γ) is the length <strong>of</strong> γ with respect <strong>to</strong> S.Pro<strong>of</strong>. Note that P n is a finite index subgroup <strong>of</strong> B n . Now the pro<strong>of</strong> followsdirectly from [21, Corollary 24].1.4 Quasi-morphismsIn this section we recall the definition <strong>and</strong> basic properties <strong>of</strong> <strong>quasi</strong>-morphisms.For more information about <strong>quasi</strong>-morphisms see [9] <strong>and</strong> [14].Definition 1.4.1. Let G be a group. A function φ : G → R is called a <strong>quasi</strong>morphismif there exists K φ ≥ 0 such that|φ(gh) − φ(g) − φ(h)| ≤ K φfor all g, h ∈ G. The smallest <strong>of</strong> such K φ is called the defect <strong>of</strong> φ <strong>and</strong> will bedenoted by D φ .Examples 1.4.2. (i) Bounded functions <strong>and</strong> homomorphisms are <strong>quasi</strong>-morphisms.(ii) Let F n be a free group on n genera<strong>to</strong>rs x 1 , . . . , x n . Brooks [12] showed thatthere are many <strong>quasi</strong>-morphisms on F n which are not homomorphisms. Namelylet ω be a reduced word in x 1 , . . . , x n <strong>and</strong> x −11 , . . . , x −1n . Define ρ ω : F n → Zby setting ρ ω (x) <strong>to</strong> be the difference between the number <strong>of</strong> times ω <strong>and</strong> ω −1occur as subwords <strong>of</strong> x when x is written as a reduced word in x 1 , . . . , x n <strong>and</strong>x −11 , . . . , x −1 . Then ρ ω is a <strong>quasi</strong>-morphism.nDefinition 1.4.3. A <strong>quasi</strong>-morphism φ is called homogeneous if for every g ∈ G<strong>and</strong> m ∈ Zφ(g m ) = mφ(g).20


Any <strong>quasi</strong>-morphism φ can be homogenized: setting˜φ(g) :=limk→+∞ φ(gk )/k (1.19)we get a homogeneous (possibly trivial) <strong>quasi</strong>-morphism ˜φ. Throughout thethesis the induced homogeneous <strong>quasi</strong>-morphism obtained by the homogenization<strong>of</strong> φ will be denoted by ˜φ.Now we recall some properties <strong>of</strong> homogeneous <strong>quasi</strong>-morphisms, see e.g.[14]:˜φ(hgh −1 ) = ˜φ(g) ∀g, h ∈ G (1.20)˜φ(gh) = ˜φ(g) + ˜φ(h) ∀g, h ∈ G such that gh = hg. (1.21)Given a group G, its first derived subgroup is the subgroup G ′ generatedby commuta<strong>to</strong>rs [g, h] = ghg −1 h −1 <strong>of</strong> elements g, h ∈ G, i.e. each element g ′ <strong>of</strong>G ′ can be written as a product ∏ ki=1 [g i, h i ]. The smallest integer k for whichsuch an expression exists is called the commuta<strong>to</strong>r length <strong>of</strong> g ′ <strong>and</strong> is denotedby comm(g ′ ). Let us define the stable commuta<strong>to</strong>r length <strong>of</strong> g ′ by setting||g ′ 1|| := limn→∞ n comm((g′ ) n ).Lemma 1.4.4 ([9]). Let φ : G → R be a non-trivial homogeneous <strong>quasi</strong>morphismwith a defect D φ ≠ 0. Then the stable commuta<strong>to</strong>r length function|| · || : G ′ → N is non-trivial.In [9], Bavard showed that if || · || is a non-trivial function, then there existsa non-trivial homogeneous <strong>quasi</strong>-morphism φ with D φ ≠ 0.1.5 Calabi homomorphismRecall that D := Diff ∞ (D 2 , ∂D 2 , area) st<strong>and</strong>s for the group <strong>of</strong> area-preservingdiffeomophisms <strong>of</strong> the unit closed two dimensional disc D 2 in the Euclideanplane which are identical near the boundary ∂D 2 . In this section we recall thedefinition <strong>and</strong> properties <strong>of</strong> the Calabi homomorphism C : D → R. Denote byω the st<strong>and</strong>ard symplectic (i.e. the area) form on D 2 ⊂ R 2 .Definition 1.5.1. For every g ∈ D <strong>and</strong> a differential 1-form λ on D 2 , such thatdλ = ω, there exists a unique function H : D 2 → R, such that H is zero near21


the boundary <strong>and</strong> dH = g ∗ λ − λ. Define C : D → R by∫C(g) = Hω.This map is well defined <strong>and</strong> it is a homomorphism, see [5, 13], cf. [25]. It iscalled the Calabi homomorphism.D 2The celebrated theorem <strong>of</strong> Banyaga [5] states that ker(C) is a simple group.It follows that every homomorphism µ : D → R is given by µ = κ ◦ C, whereκ : R → R is a homomorphism. Note that if κ is continuous then there existsa real constant K such that µ = KC.Let H : D 2 × R → R be a C ∞ -function such that H t is zero near ∂D 2 foreach t ∈ R. Here H t := H(−, t). The Hamil<strong>to</strong>nian vec<strong>to</strong>r field X Ht is definedby the following equation:dH t (−) = ω(−, X Ht ).The flow ϕ t H generated by the vec<strong>to</strong>r field X H tis called the Hamil<strong>to</strong>nian flow.For each t the diffeomorphism ϕ t H is identity near ∂D2 <strong>and</strong> (ϕ t H )∗ ω = ω. Wesay that the diffeomorphism ϕ 1 H is the diffeomorphism generated by H. Anarea-preserving diffeomorphism ϕ is called au<strong>to</strong>nomous if ϕ = ϕ 1 H for sometime-independent Hamil<strong>to</strong>nian H.Proposition 1.5.2 ([25]). Let ϕ t H be an au<strong>to</strong>nomous Hamil<strong>to</strong>nian flow generatedby H : D 2 → R. Then∫C(ϕ t H) = −2t Hω.Proposition 1.5.3 ([5, 25]). The Calabi homomorphism C is not continuousin the C 0 -<strong>to</strong>pology on D, but it is continuous in C 1 -<strong>to</strong>pology.D 222


Chapter 2Quasi-morphisms on braidgroups defined by knot<strong>invariants</strong>It is shown in [31] that the group B n admits infinitely many linearly independenthomogeneous <strong>quasi</strong>-morphisms for every integer n > 2. However none <strong>of</strong> these<strong>quasi</strong>-morphisms are constructed geometrically. In [26] Gambaudo <strong>and</strong> Ghysgave a geometric construction <strong>of</strong> a family <strong>of</strong> <strong>quasi</strong>-morphisms sign n on groupsB n defined as follows:sign n (α) := sign(̂α),where ̂α is an n-component link obtained by closing α in the st<strong>and</strong>ard way (seeFigure 1.3). In [26] Gambaudo <strong>and</strong> Ghys proved that sign n is a <strong>quasi</strong>-morphismwith a defect bounded by n − 1. In this chapter we show that knot <strong>invariants</strong><strong>of</strong> certain type define <strong>quasi</strong>-morphisms on B n in a similar way.2.1 Main TheoremOne <strong>of</strong> the ways <strong>to</strong> find a <strong>quasi</strong>-morphism on B n is <strong>to</strong> find an R−valued invariantI <strong>of</strong> (iso<strong>to</strong>py classes <strong>of</strong>) links in S 3 so that|I(̂αβ) − I(̂α) − I(̂β)| ≤ K I ,where K I ≥ 0 depends only on I.In this section we will formulate a sufficient condition for a knot invariantunder which it yields a <strong>quasi</strong>-morphism on B n .23


We will now describe certain ways <strong>of</strong> closing braids in<strong>to</strong> knots.Lemma 2.1.1. Let β ∈ B n . Then there exists a braid α β ∈ B n which satisfiesthe following properties:1. The closure <strong>of</strong> α β β is a knot.2. The closure <strong>of</strong> α β is a k-component unlink for some 1 ≤ k ≤ n.We will say that such a braid α β is a completing braid for β.Pro<strong>of</strong>. Let β ∈ B n . Then there exists a braid α such that ̂αβ is a knot <strong>and</strong>̂α is a k-component link for some 1 ≤ k ≤ n. We write α as a product <strong>of</strong>genera<strong>to</strong>rs σ ±1i <strong>and</strong> perform a sequence <strong>of</strong> crossing changes (i.e. replacing some<strong>of</strong> the genera<strong>to</strong>rs in the product by <strong>their</strong> inverses) until we are left with a braidα β such that ̂αβ = ̂α β β is a knot <strong>and</strong> ̂α β is the k-component unlink.Definition 2.1.2. Let I be a real-valued knot invariant. Let us fix some choices<strong>of</strong> completing braids α β for every β ∈ B n . Define a functionby Î(β) := I( ̂α β β).Î : B n → R,We will show that under certain conditions Î is a <strong>quasi</strong>-morphism.Remark 2.1.3. Note that our function Î : B n → R is slightly different froman analogous <strong>quasi</strong>-morphism defined by Gambaudo-Ghys. Their definition(Î(β) := I(̂β)) requires I <strong>to</strong> be a real-valued link invariant, but in our definitionwe require I <strong>to</strong> be only a real-valued knot invariant. For example the Ozsvath-Szabo τ invariant is defined only for knots. In case I is a real-valued linkinvariant defining a <strong>quasi</strong>-morphism Î, the <strong>quasi</strong>-morphisms Î <strong>and</strong> Î differ by aconstant which depends only on n, hence <strong>their</strong> homogenizations are equal.Theorem 2.1.4. Suppose that a real-valued knot invariant I defines a homomorphismI : Conc(S 3 ) → R,such that |I(K)| ≤ cg 4 (K), where c is a real positive constant independent <strong>of</strong>K. Then Î is a <strong>quasi</strong>-morphism on B n. Moreover, for a different sets <strong>of</strong> choices<strong>of</strong> completing braids α β for every β ∈ B n we get a (possibly different) <strong>quasi</strong>morphismon B n such that the absolute value <strong>of</strong> its difference with Î is boundedfrom above by a constant depending only on n <strong>and</strong> therefore the homogenizations<strong>of</strong> the two <strong>quasi</strong>-morphisms are equal.24


Pro<strong>of</strong>. Take any β, γ ∈ B n <strong>and</strong> let α β , α γ , α βγ be the chosen completing braidsfor β, γ <strong>and</strong> βγ.Lemma 2.1.5. There exists a cobordism S between the knots(− ̂α βγ βγ) ∗ #( ̂α β β#̂α γ γ)<strong>and</strong>such that χ(S) ≥ −6n.(− ̂α βγ βγ) ∗ # ̂α βγ βγPro<strong>of</strong>. If T is a cobordism between two links L 1 <strong>and</strong> L 2 , we will write L 1 T ∼ L 2 .By an observation <strong>of</strong> Baader (see [4], Section 4) for any braids µ, ν ∈ B n̂µνT∼̂µ ⊔ ̂ν,where χ(T ) = −n. Also note that since ̂α βγ , ̂α β , ̂α γ are unlinks with no morethan n components, we haveŜα1βγ ∼ ̂αβ ⊔ ̂α γ ,where χ(S 1 ) ≥ 1 − 2n.Thereforêα βγ βγ S 2∼ ̂α βγ ⊔ ̂β ⊔ ̂γ S 3∼S∼3̂αβ ⊔ ̂α γ ⊔ ̂β ⊔ ̂γ S 4∼ ̂α β β ⊔ ̂α γ γ S 5∼ ̂α β β#̂α γ γ,where χ(S 2 ) = −2n, χ(S 3 ) ≥ 1 − 2n (the cobordism S 3 is the disjoint union <strong>of</strong>S 1 <strong>and</strong> the trivial cobordism over ̂β ⊔ ̂γ given by a disjoint union <strong>of</strong> cylinders),χ(S 4 ) = −2n <strong>and</strong> χ(S 5 ) = −1 (we take S 5 as a saddle cobordism between thedisjoint union <strong>of</strong> two knots <strong>and</strong> <strong>their</strong> connected sum). Thuŝα βγ βγ S 6∼̂α β β#̂α γ γ<strong>and</strong> hence(− ̂α βγ βγ) ∗ # ̂α βγ βγ S 7∼(− ̂α βγ βγ) ∗ #( ̂α β β#̂α γ γ),whereχ(S 6 ) = χ(S 7 ) = χ(S 2 ) + χ(S 3 ) + χ(S 4 ) + χ(S 5 ) ≥ −6n,as required.25


Let us now finish the pro<strong>of</strong> <strong>of</strong> the theorem. Lemma 1.1.2 implies that theknot(− ̂α βγ βγ ∗ )# ̂α βγ βγis a slice knot. Therefore using Lemma 2.1.5 we getg 4 (−( ̂α βγ βγ ∗ )#( ̂α β β#̂α γ γ)) ≤ 3n,where g 4 is the four-ball genus. We know thatfor every knot K. This yieldsApplying the equalitieswe get|I(K)| ≤ cg 4 (K) (2.1)|I(−( ̂α βγ βγ ∗ )#( ̂α β β#̂α γ γ))| ≤ c · 3n.I(−K ∗ ) = −I(K), I(K 1 #K 2 ) = I(K 1 ) + I(K 2 ), (2.2)|Î(βγ) − Î(β) − Î(γ)| ≤ 3cn,which means that Î is a <strong>quasi</strong>-morphism.Finally note that for any two choices α β <strong>and</strong> α β ′ <strong>of</strong> completing braids forβ ∈ B n one can show, similarly <strong>to</strong> the pro<strong>of</strong> <strong>of</strong> Lemma 2.1.5, thatg 4 (( ̂α β β)# − ( ̂α ′ β β)∗ ) ≤ c 1 n,for some positive constant c 1 independent <strong>of</strong> β, α β , α β ′ . Combining this withequations (2.2), (2.1) we get|I( ̂α β β) − I( ̂α ′ β β)| = |I(( ̂α β β)# − ( ̂α ′ β β)∗ )| ≤ c · c 1 n.Thus different choices <strong>of</strong> completing braids yield (possibly different) <strong>quasi</strong>morphismson B n whose difference is bounded in absolute value from aboveby a constant depending only on n <strong>and</strong> therefore <strong>their</strong> homogenizations areequal.The induced homogeneous <strong>quasi</strong>-morphism is denoted by Ĩ.26


Remark 2.1.6. Note that g 4 defines a norm on Conc(S 3 ) <strong>and</strong> Theorem 2.1.4can be reformulated as follows: each element <strong>of</strong> Hom(Conc(S 3 ), R), which isLipshitz with respect <strong>to</strong> the norm defines a <strong>quasi</strong>-morphism on B n .In the next section we will present the following knot/link <strong>invariants</strong>: theOzsvath-Szabo knot invariant τ, the Rasmussen link invariant s <strong>and</strong> the linksignatures. It is well known that all <strong>of</strong> them satisfy the conditions <strong>of</strong> Theorem2.1.4. Thus by Theorem 2.1.4 they define <strong>quasi</strong>-morphisms on B n .2.2 Examples <strong>and</strong> <strong>applications</strong>2.2.1 Rasmussen <strong>quasi</strong>-morphismIn his paper [4] Baader proved that the Rasmussen link invariant s induces a<strong>quasi</strong>-morphism on the braid group B n , with a defect bounded by n + 1. Theinduced <strong>quasi</strong>-morphism was denoted by s. Alternatively, we see that s satisfiesproperties (1.1), (1.2), (1.3) <strong>and</strong> (1.4). Thus our Theorem 2.1.4 proves that sdefines a <strong>quasi</strong>-morphism on B n . The induced <strong>quasi</strong>-morphism (our definition)is denoted by ŝ. Following [26] we denote by lk : B n → Z the homomorphismfrom B n <strong>to</strong> Z by setting lk(σ ±1i ) = ±1.Theorem 2.2.1. Suppose that φ : B n → R is a <strong>quasi</strong>-morphism defined by areal-valued link invariant φ, which satisfies property (1.8), where s is substitutedby φ. Then ˜φ = lk.Pro<strong>of</strong>. Take β ∈ B n . Then for all p > 0 there exists a braid α p in B n , such that̂α p β p is a knot, l(α p ) ≤ M(n) <strong>and</strong> |lk(α p )| ≤ M(n), where l is a braid length<strong>and</strong> M(n) is some real positive constant which depends only on n. For exampleif β ∈ P n , than we can takeα p = α = σ 1 · . . . · σ n−1for all p ∈ N, see Figure 2.1. Note thatw(D(̂α p β p )) = lk(α p β p ) <strong>and</strong> o(D(̂α p β p )) = n.It follows from the property (1.8) that1 + lk(α p β p ) − n ≤ φ(α p β p ) ≤ −1 + lk(α p β p ) + n.27


z1z2z3znzz1 2z nFigure 2.1: The braid α = σ 1 · . . . · σ n−1For all p > 0 there exist a constant K(n) > 0, such that |φ(α p )| ≤ K(n). Thus:1limp→∞ p (1 + lk(α pβ p 1) − n) ≤ limp→∞ p φ(α pβ p 1) ≤ limp→∞ p (−1 + lk(α pβ p ) + n).1Therefore lim φ(αp→∞ p pβ p ) = lk(β). Now φ is a <strong>quasi</strong>-morphism with a defect D φ<strong>and</strong> so|φ(α p β p ) − φ(β p )| ≤ |φ(α p )| + D φ < K(n) + D φ .It follows thatlk(β) = limp→∞φ(α p β p )p= limp→∞φ(β p )p= ˜φ(β).Definition 2.2.2. A link diagram is called alternating if the crossings alternateunder, over, under, over, <strong>and</strong> so on as one travels along each component <strong>of</strong> thelink. A link is called alternating if it has an alternating diagram. A braidα ∈ B n is called alternating if ̂α is an alternating link.We denote by ŝign n ( ˜sign n ) the (homogeneous) <strong>quasi</strong>-morphism on B n inducedby the link invariant sign.Corollary 2.2.3. Each α ∈ B n satisfies ˜s(α) = lk(α). If γ ∈ B n is an alternatingbraid, then˜sign n (γ) = lk(γ).28


Pro<strong>of</strong>. The first statement follows directly from Theorem 2.2.1.Now we will prove the second statement. For each p ∈ N the braid γ p isalternating <strong>and</strong> there exists an alternating braid α p such that l(α p ) < M(n),where ̂α p γ p is an alternating knot <strong>and</strong> M(n) is some real positive constant whichdepends only on n. It follows from property (1.6) thatŝignlim n (α p γ p )p→∞ p= limp→∞ŝ(α p γ p )p= limp→∞lk(α p γ p )p= lk(γ).Using the fact that ŝign n is a <strong>quasi</strong>-morphism we get the following equalityŝignlim n (α p γ p )p→∞ p= ˜sign n (γ).In [4] Baader defined the following <strong>quasi</strong>-morphism:s − lk + n − 1 : B n → R,where s(β) := s(̂β) (Gambaudo-Ghys definition). This <strong>quasi</strong>-morphism mapsall positive braids <strong>to</strong> zero. Therefore it descends <strong>to</strong> a <strong>quasi</strong>-morphism onB n /⟨∆ n ⟩, where ∆ 2 n is the positive genera<strong>to</strong>r <strong>of</strong> the center Z(B n ) <strong>of</strong> B n . In[4] Baader asked whether this <strong>quasi</strong>-morphism is bounded for n > 2. Here wegive a positive answer <strong>to</strong> this question.Proposition 2.2.4. The <strong>quasi</strong>-morphism s − lk + n − 1 : B n → R is bounded.Pro<strong>of</strong>. Take any braid β ∈ B n <strong>and</strong> a braid α β , such that ̂α β is an unlink. Usingthe fact that s is a <strong>quasi</strong>-morphism with a defect bounded by n + 1, <strong>and</strong> thats(α β ) ≤ n − 1 <strong>and</strong> lk(α β ) ≤ n − 1 we get the following inequality:|s(β) − lk(β)| ≤ |s(α β β) − lk(α β β)| + 3n − 1.It follows from property (1.8) thatThis yields|s(α β β) − lk(α β β)| ≤ n − 1.|s(β) − lk(β) + n − 1| ≤ 5n − 3.The last inequality shows that the <strong>quasi</strong>-morphism s − lk + n − 1 is boundedon B n <strong>and</strong> thus also on B n /⟨∆ n ⟩.29


2.2.2 Connection between τ, s <strong>and</strong> braid numberThe <strong>invariants</strong> s <strong>and</strong> 2τ share similar properties <strong>and</strong> coincide on positive <strong>and</strong>alternating knots. It was conjectured by Rasmussen [56] that they are equal.This conjecture was disproved by Hedden <strong>and</strong> Ording [32]. In this subsectionwe show that the difference between 2τ <strong>and</strong> s is bounded by the braid number<strong>and</strong> that ˜s = 2˜τ.Let L be an oriented link in S 3 <strong>and</strong> D(L) be any diagram <strong>of</strong> L. Recall thatin Section 1.1 we associated, <strong>to</strong> each such D(L), a family <strong>of</strong> Seifert circles.Lemma 2.2.5. Let K be an oriented knot in S 3 <strong>and</strong> D(K) be any diagram <strong>of</strong>K. Then1 + w(D(K)) − o(D(K)) ≤ 2τ(K) ≤ −1 + w(D(K)) + o(D(K)), (2.3)where o(D(K)) is the number <strong>of</strong> Seifert circles <strong>of</strong> D(K).Pro<strong>of</strong>. The pro<strong>of</strong> <strong>of</strong> the lower bound in (2.3) is exactly the same as in [59],where s is substituted by 2τ. We present it for the reader’s convenience. Recallthat0 ≤ 2τ(K + ) − 2τ(K − ) ≤ 2, (2.4)τ(K ∗ ) = −τ(K). (2.5)It follows from property (1.14) that the equation 2τ(K 1 ) = s(K 1 ) is true forany positive knot K 1 . Thus for every positive knot diagram D(K 1 ) the followingequation holds:2τ(K 1 ) = w(D(K 1 )) − o(D(K 1 )) + 1. (2.6)It means that the left inequality in (2.3) is then an equality. When a positivecrossing <strong>of</strong> D(K) is changed in<strong>to</strong> a negative one, the number 1 + w(D(K)) −o(D(K)) decreases by 2, while 2τ(K) decreases by at most 2, because <strong>of</strong> (2.4).Hence, the left inequality in (2.3) is preserved. The upper bound in (2.3) istrue, because <strong>of</strong> (2.5) combined with the lower bound.Proposition 2.2.6. The knot invariant τ defines a <strong>quasi</strong>-morphism on B n <strong>and</strong>2˜τ = lk.Pro<strong>of</strong>. The first statement follows directly from Theorem 2.1.4. The secondstatement follows from Lemma 2.2.5 <strong>and</strong> from the pro<strong>of</strong> <strong>of</strong> Theorem 2.2.1.30


Recall that every knot K in R 3 can be presented as a closure <strong>of</strong> some braidin B n . The braid number <strong>of</strong> K is the minimal such n. It is denoted by br(K).Theorem 2.2.7. Let ŝ <strong>and</strong> ̂τ be the <strong>quasi</strong>-morphisms on B n which are inducedfrom Rasmussen <strong>and</strong> Ozsvath-Szabo knot <strong>invariants</strong>. Then for every β ∈ B nthe following inequality holds |ŝ(β) − 2̂τ(β)| ≤ 2(n − 1).Pro<strong>of</strong>. As we explained before both s <strong>and</strong> 2τ satisfy property (1.8). It meansthat|ŝ(β) − lk(α β β)| ≤ n − 1 <strong>and</strong> |2̂τ(β) − lk(α β β)| ≤ n − 1.Thus by the triangle inequality we have|ŝ(β) − 2̂τ(β)| ≤ 2(n − 1).Corollary 2.2.8. For every knot K the following inequality holds:|s(K) − 2τ(K)| ≤ 2(br(K) − 1).2.2.3 Signature <strong>quasi</strong>-morphismsQuasi-morphism defined by the classical signature link invariantRecall that the signature link invariant was denoted by sign <strong>and</strong> the inducedhomogeneous <strong>quasi</strong>-morphism on B n was denoted by ˜sign n .Proposition 2.2.9. 1. For every n > 2 the homogeneous <strong>quasi</strong>-morphism˜sign n : B n → R is not a homomorphism.2. The restriction <strong>of</strong> ˜sign n <strong>to</strong> P n is not a homomorphism.Pro<strong>of</strong>. Claim 1. Take a braid α n ∈ B n for n > 2, where α n = σ 1 · . . . · σ n−1 .Then it follows from [47, page 148] that˜sign n (α n ) = n 2 , if n is even, <strong>and</strong> ˜sign n (α n ) = n2 − 1, if n is odd.2nIt follows that if n is even thenn2 = ˜sign n (α n ) ≠ ˜sign n (α n−1 ) + ˜sign n (σ n−1 ) = (n − 1)2 − 1+ 1,2n − 231


<strong>and</strong> if n is odd thenn 2 − 12n= ˜sign n (α n ) ≠ ˜sign n (α n−1 ) + ˜sign n (σ n−1 ) = n − 12+ 1.The pro<strong>of</strong> <strong>of</strong> Claim 2 will be given in the next chapter.Let η i,n be a pure braid in P n , in which the i−th str<strong>and</strong> makes one twist inthe positive direction around previous i − 1 str<strong>and</strong>s, <strong>and</strong> η − i,n is a pure braid inP n , in which the i−th str<strong>and</strong> makes one twist in the negative direction aroundprevious i − 1 str<strong>and</strong>s, see Figure 2.2. Note that all <strong>of</strong> these braids commutewith each other. These braids will play a crucial role in the next chapter. Ini−1iFigure 2.2: Braid η i,n[26] Gambaudo <strong>and</strong> Ghys computed ˜sign n (η i,n ). We present this computationin detail.Lemma 2.2.10 ([26]). For all i, n ∈ N we have{i, if i is even,˜sign n (η i,n ) =i − 1, if i is odd.Pro<strong>of</strong>. The <strong>to</strong>rus link K(p, q) has a braid representation (σ 1 · . . . · σ p−1 ) q . NotethatK(n, np) = η p 2,n · . . . · η p n,n = (η 2,n · . . . · η n,n ) p .It follows that˜sign n (K(n, n)) = ˜sign n (η 2,n · . . . · η n,n ) =32n∑i=2˜sign n (η i,n ).


Also note that for all k, r ∈ N, such that k ≥ i <strong>and</strong> r ≥ i, the following equationholds˜sign n (η i,k ) = ˜sign n (η i,r ).It follows that˜sign n (η i,n ) = ˜sign n (K(i, i)) − ˜sign n (K(i − 1, i − 1)).It follows from [47, Theorem 7.5.1] that˜sign n (K(i, i)) = i2 − 1, if i is odd,2˜sign n (K(i, i)) = i2 2 ,if i is even.It follows that˜sign n (η i,n ) = i2 − 12−(i − 1)2, if i is odd,2Thus we get˜sign n (η i,n ) = i2 2 − (i − 1)2 − 1, if i is even.2{i, if i is even,˜sign n (η i,n ) =i − 1, if i is odd.ω-signature <strong>quasi</strong>-morphismsGambaudo <strong>and</strong> Ghys proved in [26] that sign ω induces a <strong>quasi</strong>-morphism onB n . Alternatively, Theorem 2.1.4 shows that sign ω induces a <strong>quasi</strong>-morphismon B n . The homogeneous <strong>quasi</strong>-morphism will be denoted as usual by ˜sign ω .In [27] Gambaudo <strong>and</strong> Ghys proved the following proposition.Proposition 2.2.11 ([27], Proposition 5.2). Let ω = e 2πiθ , whereθ ∈ [0, 1] ∩ Q. Then˜sign ω (σ 1 · . . . · σ n−1 ) = 2(n − 2l + 1)θ +for l−1n ≤ θ ≤ l , where 1 ≤ l ≤ n. n332l(l − 1)n


It follows from Proposition 5.2 in [27] thatfor l−1 ≤ θ ≤ l , where 1 ≤ l ≤ n.n n˜sign ω (K(n, n)) = 2n(n − 2l + 1)θ + 2l(l − 1) (2.7)Proposition 2.2.12. Let ω = e 2πiθ , where θ ∈ [0, 1] ∩ Q. Then:⎧4(i − 1)θ, if 0 ≤ θ ≤ 1 i ,˜sign ω (η i,n ) =⎪⎨4(l − 1)(1 − θ), if l − 1i≤ θ ≤ l − 1i − 1 , 2 ≤ l ≤ i − 1,4(i − 1)θ, if l − 1i − 1 ≤ θ ≤ l i , 2 ≤ l ≤ i − 1,⎪⎩4(i − 1)θ, if i − 1i≤ θ ≤ 1.(2.8)Pro<strong>of</strong>. Note that˜sign ω (η i,n ) = ˜sign ω (K(i, i)) − ˜sign ω (K(i − 1, i − 1)).After a simple calculation this equality <strong>and</strong> (2.7) yield the result.Denote byAP n := ⟨η 2,n , η − 2,n, . . . , η n,n , η − n,n⟩the subgroup <strong>of</strong> P n generated by η ±1i,n , <strong>and</strong> defineV AP = {˜φ| APn : AP n → R where ˜φ is a homogeneous <strong>quasi</strong>-morphism on B n }.Proposition 2.2.13. Denote by ω(θ) = ω 2πiθ . ThenB ={˜signω ( 1 2) , 2 ˜sign ω ( 1 2) − 3 ˜sign ω ( 1 3) , . . . , (n − 1) ˜sign ω ( 1n−1) − n ˜sign ω ( 1 n)is a basis for V AP .}34


Pro<strong>of</strong>. If n = 2, then ˜sign ω( 1 2) (η 2,2) = 2, <strong>and</strong> the pro<strong>of</strong> follows. Let n > 2. For1 ≤ i ≤ n − 1, the equality()(n − 1) ˜sign ω( n−1) − n ˜sign 1ω((ηn)1 i,n ) = 0follows from (2.8). We also have the following equality:()(n − 1) ˜sign ω ( n−1) − n ˜sign 1ω ((ηn)1 n,n ) = −4.Note also that ˜sign ω (η i,i ) = ˜sign ω (η i,n ). For each 3 ≤ i ≤ n <strong>and</strong> for each2 ≤ j ≤ i − 1 this yields()(i − 1) ˜sign ω( i−1) − i ˜sign 1ω( 1 i )(η j,i ) = 0,<strong>and</strong> ()(i − 1) ˜sign ω ( i−1) − i ˜sign 1ω ( 1 i )(η i,i ) = −4.Thus B is a basis <strong>of</strong> V AP .35


Chapter 3Generalized Gambaudo-GhysconstructionIn this chapter we present a generalization <strong>of</strong> the Gambaudo-Ghys construction<strong>of</strong> a family <strong>of</strong> <strong>quasi</strong>-morphisms on the group D := Diff ∞ (D 2 , ∂D 2 , area). Wepresent properties <strong>of</strong> these <strong>quasi</strong>-morphisms. Further we investigate the <strong>quasi</strong>morphismson D that are induced by s, τ <strong>and</strong> sign.3.1 Gambaudo-Ghys constructionDenote the space <strong>of</strong> Hamil<strong>to</strong>nians H : D 2 × [0, 1] → R, such that the support <strong>of</strong>H t := H(·, t) is compact for any t, by H. We denote the space <strong>of</strong> au<strong>to</strong>nomous(i.e. time-independent) Hamil<strong>to</strong>nians by H 0 .The following construction, essentially contained in Gambaudo-Ghys [26],takes a homogeneous <strong>quasi</strong>-morphism on P n <strong>and</strong> produces from it a <strong>quasi</strong>morphismon D.Let x = (x 1 , . . . , x n ) be any point in X n . Take g ∈ D <strong>and</strong> any path g t ,0 ≤ t ≤ 1, in D between Id <strong>and</strong> g. Connect z <strong>to</strong> x by a straight line in (D 2 ) n ,then act on x with the path g t , <strong>and</strong> then connect g(x) <strong>to</strong> z by the straight linein (D 2 ) n . We get a loop in (D 2 ) n . More specifically it looks as follows. Connectz i <strong>to</strong> x i by straight lines l 1,i : [ 0, 3] 1 → D 2 in the disc, then act with the path1g 3t−1 , ≤ t ≤ 2, on each x 3 3 i, <strong>and</strong> then connect g(x i ) <strong>to</strong> z i by straight linesl 2,i : [ 2, 1] → D 2 in the disc, for all 1 ≤ i ≤ n. It is a well known fact that D3coincides with the group <strong>of</strong> compactly supported Hamil<strong>to</strong>nian diffeomorphisms<strong>of</strong> D 2 (see e.g. [45]). Thus any identity-based path {h t } in D, 0 ≤ t ≤ 1, can36


e viewed as a Hamil<strong>to</strong>nian flow generated by a (time-dependent) Hamil<strong>to</strong>nianH : D 2 × [0, 1] → R, such that the support <strong>of</strong> H t := H(·, t) is compact for anyt. It follows that for almost all n-tuples <strong>of</strong> different points x 1 , . . . , x n in thedisc the concatenations <strong>of</strong> the paths l 1,i : [ 0, 3] 1 → D 2 , g 3t−1 : [ 1, 2 3 3]→ D 2 <strong>and</strong>l 2,i : [ 2, 1] → D 2 , i = 1, . . . , n, yield a loop in X3 n . The homo<strong>to</strong>py type <strong>of</strong> thisloop is an element in P n , which is independent <strong>of</strong> the choice <strong>of</strong> g t because D iscontractible [23]. We will denote this element by γ(g; x).Let ˜φ be a homogeneous <strong>quasi</strong>-morphism on P n . Set∫Φ(g) = ˜φ(γ(g; x))dx, (3.1)X nwhere dx = dx 1 · . . . · dx n .Lemma 3.1.1 (c.f. [26]). The function Φ is well defined. It is a <strong>quasi</strong>-morphismon D.As an immediate corollary we get that the formula∫1˜Φ(g) := lim ˜φ(γ(g p ; x))dxp→∞ pyields a homogeneous <strong>quasi</strong>-morphism ˜Φ : D → R.X nPro<strong>of</strong>. Step 1. Let us prove that |Φ(h)| < ∞ for h ∈ D.For any iso<strong>to</strong>py h t , 0 ≤ t ≤ 1, in D between h = h 1 <strong>and</strong> Id, any x ∈ X n <strong>and</strong>any 1 ≤ i, j ≤ n, i ≠ j denoteDenote byl i,j := h t(x i ) − h t (x j )∥h t (x i ) − h t (x j )∥ : [0, 1] → S1 .L i,j (x) = 1 ∫ 12π0∂∥∂t (l i,j)∥ dt, (3.2)where ∥ · ∥ is the Euclidean norm. Note that L i,j (x) is the length <strong>of</strong> the pathl i,j divided by 2π. Thus we get the following inequalityn∑2 (L i,j (x) + 4) ≥ l(γ(h; x)). (3.3)i


Take any finite generating set S <strong>of</strong> P n (there exists at least one). Note that forany homogeneous <strong>quasi</strong>-morphism ˜φ : P n → R one has()|˜φ(γ)| ≤ D ˜φ + max |˜φ(ξ)| l S (γ), (3.4)ξ∈Swhere l S (γ) is the length <strong>of</strong> a word γ with respect <strong>to</strong> S. Recall that l(γ) is thelength <strong>of</strong> γ with respect <strong>to</strong> the set {σ i } n−1i=1 . It follows from Lemma 1.3.5 thatthere exist two positive constants K 1,S <strong>and</strong> K 2,S , which are independent <strong>of</strong> γ,such thatl S (γ) ≤ K 1,S · l(γ) + K 2,S .It follows from (3.4) that|˜φ(γ(h; x))| ≤ N 1 l(γ(h; x)) + N 2 , (3.5)where N 1 = K 1,S (D ˜φ + max |˜φ(ξ)|) <strong>and</strong> N 2 = K 2,S (D ˜φ + max |˜φ(ξ)|). Inequalities(3.3) <strong>and</strong> (3.5) yield the followingξ∈S ξ∈Sinequality:It follows that|˜φ(γ(h; x))| ≤ N 1|Φ(h 1 )| ≤ 2N 1 · vol((D 2 ) n−2 )n∑i


∥Let H ∈ H be a Hamil<strong>to</strong>nian generating the flow h t , 0 ≤ t ≤ 1. Note that∂t∥ = ∥(∇H t )(h t (x))∥ ≤ max ∥(∇H t )(x)∥. Denotex∈D 2 ,t∈[0,1]∥ ∂h t(x)ThusM H :=∥ ∥∥∥ ∂∂t (l i,j)∥ ≤max ∥(∇H t )(x)∥ .x∈D 2 ,t∈[0,1]4M H∥h t (x i ) − h t (x j )∥ .Recall that h t is a Hamil<strong>to</strong>nian flow <strong>and</strong> therefore is area-preserving. Then⎛⎞∫ ∫11∂2π ∥∂t (l i,j)∥ dtdx idx j ≤ 4M ∫H⎝12π ∥x i − x j ∥ dx idx j⎠ .D 2 ×D 2 0D 2 ×D 2Now we show that∫D 2 ×D 2 1∥x i − x j ∥ dx idx jis well defined. It follows by Fubini-Tonelli theorem that∫1∥x i − x j ∥ dx idx j ≤D 2 ×D 2∫≤D 2⎛⎜⎝∫B δ (x j )1∥x i − x j ∥ dx i +∫D 2 \(D 2 ∩B δ (x j ))⎞1∥x i − x j ∥ dx ⎟i⎠ dx j ,where δ ≤ 1 is some positive constant. By using the polar coordinates on B 2 δ(x j )we get∫∫) (1∥x i − x j ∥ dx idx j < 2πδ + vol(D2 )dx j = π 2 2δ + 1 ).δδD 2 ×D 2DenoteN ′ := 4N 1 n(n + 1)⎛D 2 (⎝ vol((D2 ) n−2 )2π⎛⎝∫D 2 ×D 2 139⎞⎞∥x i − x j ∥ dx idx j⎠ + vol((D 2 ) n ) ⎠ ,


then|Φ(h)| ≤ N ′ · M H + A(X n ).Step 2. Let us prove that Φ is a <strong>quasi</strong>-morphism. Let g 1 <strong>and</strong> g 2 be elements<strong>of</strong> D. Note that for almost all x ∈ X nγ(g 1 g 2 ; x) = γ(g 1 ; g 2 (x)) · γ(g 2 ; x).The diffeomorphisms g 1 <strong>and</strong> g 2 are area-preserving. Thus∫=∣It follows thatX n|Φ(g 1 g 2 ) − Φ(g 1 ) − Φ(g 2 )| =˜φ n (γ(g 1 g 2 ; x)) − ˜φ n (γ(g 1 ; g 2 (x))) − ˜φ n (γ(g 2 ; x))dx∣ .∫|Φ(g 1 g 2 ) − Φ(g 1 ) − Φ(g 2 )| ≤ D ˜φ .X nFrom now on the homogeneous <strong>quasi</strong>-morphism on D induced by a homogeneous<strong>quasi</strong>-morphisms ˜φ n : P n → R will be denoted by ˜Φ n .3.2 Reeb graphs <strong>and</strong> computations forau<strong>to</strong>nomous diffeomorphismsIn this section we define the notion <strong>of</strong> a Morse-type Hamil<strong>to</strong>nian. We also recallthe notion <strong>of</strong> a Reeb graph. Recall that the space <strong>of</strong> au<strong>to</strong>nomous Hamil<strong>to</strong>nians isdenoted by H 0 . It follows from [46, Theorem 2.7] that Morse-type Hamil<strong>to</strong>niansform a C 1 -dense subset <strong>of</strong> H 0 . For any Hamil<strong>to</strong>nian H in this subset we presenta calculation <strong>of</strong> ˜Φ on the time-one flow <strong>of</strong> H (Theorem 3.2.4). Our result ispresented in terms <strong>of</strong> the integral <strong>of</strong> the push-forward <strong>of</strong> H <strong>to</strong> its Reeb graphagainst a certain signed measure on the graph. In the end <strong>of</strong> this section weshow that ˜Φ is continuous in C 1 -<strong>to</strong>pology on the space H 0 (see Theorem 3.3.5).Definition 3.2.1. We say that a function H ∈ H 0 is <strong>of</strong> Morse-type if:1. There exists a connected open neighborhood U <strong>of</strong> ∂D 2 , such that ∂U \ ∂D 2is a smooth simple curve, H| U ≡ 0 <strong>and</strong> H has no degenerate critical points in40


D 2 \ U.2. There exists a connected open neighborhood V <strong>of</strong> ∂D 2 , such that V ⊃ U<strong>and</strong> H| V \U has no critical points.3. The inequality H(x) ≠ H(y) holds for each two non-degenerate differentcritical points x <strong>and</strong> y.Now let us explain how we associate <strong>to</strong> H <strong>and</strong> ˜φ the Reeb graph T <strong>of</strong> H<strong>and</strong> the signed measure µ on T .3.2.1 Angle-action coordinates <strong>and</strong> the Reeb graphDefinition 3.2.2. A charged tree T is a finite tree equipped with a signedmeasure, such that each edge has a <strong>to</strong>tal finite measure.Now we are ready <strong>to</strong> explain how we associate <strong>to</strong> (H, ˜φ) a charged tree(T, µ). Let H ∈ H 0 be a Morse-type function with l critical points in D 2 \ U,where U is as in Definition 3.2.1. Let c = ∂U \ ∂D 2 <strong>and</strong> suppose that the area<strong>of</strong> the domain U is 2πε for some ε > 0. Let h 1 be a time-one flow generated byH.Step 1. Here we recall the notion <strong>of</strong> angle-action symplectic coordinates.We remove from D 2 all singular level curves inside D 2 \ U, the curve c<strong>and</strong> ∂D 2 . We get l open annuli A i (a punctured disc is also viewed as anannulus) <strong>and</strong> an open annulus A l+1 = U ◦ . Each D 2 \ A i has two connectedcomponents. We denote by CA i the component which does not contain ∂D 2 ,<strong>and</strong> by a i := ∂(CA i ) ∩ ∂A i . By Liouville-Arnold theorem (see [3]) on eachone <strong>of</strong> the annuli A i (1 ≤ i ≤ l), there exist so-called angle-action symplecticcoordinates (J i , θ i ), θ i ∈ [0, 2π], <strong>and</strong> a C ∞ -function[ area(CAi ) i :, area(CA ]i) + area(A i )→ R,2π2πsuch that on each level curve δ i in A i one has H| δi = i (J i ), where the coordinateJ i along δ i is equal <strong>to</strong> the sum <strong>of</strong> area(CA i)<strong>and</strong> the area <strong>of</strong> the annulus bounded2πby δ i <strong>and</strong> a i , divided by 2π. In these coordinates h 1 moves points on each levelcurve with a constant speed ′ i := ∂ i∂J i.41


Take the coordinates (J l+1 , θ l+1 ) on A l+1 , where J l+1 ∈ [ 12 − ε, 1 2]is the areacoordinate divided by 2π, <strong>and</strong> θ l+1 is the angle coordinate. Denote byNote that l+1 = 0, <strong>and</strong>∫D 2 H(p, q)dpdq =R i := area(A i)2πl∑∫i=1A iCR i := area(CA i).2πH(p, q)dpdq = 2πl∑i=1CR∫i +R iCR i i (J i )dJ i ,where p, q are the coordinates on R 2 .Step 2. Let H : D 2 → R be a Morse-type function with l critical points inD 2 \ U. Define the equivalence relation ∼ on points on D 2 by x ∼ y wheneverx, y are in the same connected component <strong>of</strong> a level set <strong>of</strong> H. The Reeb graphT corresponding <strong>to</strong> H is the quotient <strong>of</strong> D 2 by the relation ∼. The edges <strong>of</strong>T come from the annuli in D 2 fibered by level loops, the valency-one verticescorrespond <strong>to</strong> the min/max critical points <strong>of</strong> H <strong>and</strong> <strong>to</strong> ∂D 2 , <strong>and</strong> the valencythreevertices correspond <strong>to</strong> the saddle critical points <strong>of</strong> H. In our case T is arooted tree with the root being the vertex <strong>of</strong> T corresponding <strong>to</strong> the domainU ⊃ ∂D 2 . Each edge e j in T corresponds <strong>to</strong> an annulus A j for each 0 ≤ j ≤ l.Example 3.2.3. Figure 3.1 shows a Reeb graph with 5 edges <strong>and</strong> 6 vertices,such that 3 <strong>of</strong> them correspond <strong>to</strong> the maximum points, 2 <strong>of</strong> them correspond<strong>to</strong> the saddle points <strong>and</strong> the root corresponds <strong>to</strong> the region bounded by thecurves c <strong>and</strong> ∂D 2 .We call a Reeb graph simple if it has only 2 vertices <strong>and</strong> one edge. Notethat if H = H(p 2 + q 2 ) is a mono<strong>to</strong>ne function <strong>of</strong> p 2 + q 2 such that 0 ∈ D 2 isits unique non-degenerate critical point, then T is simple.Step 3. The action coordinate on any <strong>of</strong> the annuli induces a coordinate J jon the corresponding edge e j <strong>of</strong> T . Thus each H ∈ H 0 descends <strong>to</strong> a function : T → R, where ′ = ′ j on each open edge e j . Here j ∈ {0, . . . , l}. For ahomogeneous <strong>quasi</strong>-morphism ˜φ : P n → R we define a signed measure µ on Tby settingn∑( ) ( n−i n 1dµ(J j ) := (2π) n ˜φ(η i,n )i (J j ) j) i−1i 2 − J dJ j ,i=2for each 0 ≤ j ≤ l, where J j is the coordinate on each edge e j .42


AA1 2A 1 A 2A4AA3A54A 3A 5curve cFigure 3.1: Reeb graph which corresponds <strong>to</strong> the function H with 3 maximumpoints <strong>and</strong> 2 saddle points3.2.2 Main TheoremWe will work with homogeneous <strong>quasi</strong>-morphisms on P n which are restrictions<strong>of</strong> homogeneous <strong>quasi</strong>-morphisms on B n , or, equivalently (see e.g. [31]), withhomogeneous <strong>quasi</strong>-morphisms ˜φ : P n → R such that ˜φ(h) = ˜φ(g −1 hg) foreach g ∈ B n <strong>and</strong> each h ∈ P n . Such a <strong>quasi</strong>-morphism on P n will be calledB n -invariant. This is a technical condition−the results below are likely <strong>to</strong> holdfor all homogeneous <strong>quasi</strong>-morphisms on P n . Now we will state <strong>and</strong> prove ourmain theorem in this chapter.Theorem 3.2.4. Let H ∈ H 0 be a Morse-type function, ˜φ : P n → R a B n -invariant homogeneous <strong>quasi</strong>-morphism, <strong>and</strong> ˜Φ the corresponding homogeneous<strong>quasi</strong>-morphism on D. Then∫˜Φ(h 1 ) = ′ dµ, (3.8)where h 1 is the time-one Hamil<strong>to</strong>nian flow generated by H.T43


Pro<strong>of</strong>. Step 1. Suppose that the corresponding Reeb graph T has l labelededges. Thus T is a rooted labeled tree. Recall that A l+1 corresponds <strong>to</strong> theroot <strong>of</strong> T .Take all n-tuples <strong>of</strong> points x = (x 1 , . . . , x n ) ∈ X n such that for each x i , x j ∈D 2 \ A l+1 we require that x i , x j are not critical points <strong>of</strong> H, H(x i ) ≠ H(x j ) forall i ≠ j, <strong>and</strong> i < j if the corresponding points <strong>to</strong> x i <strong>and</strong> x j in T lie on the sameedge such that the corresponding action coordinates J i <strong>and</strong> J j satisfy J i < J j .For each x i , x j ∈ A l+1 we require that the corresponding action coordinates J i<strong>and</strong> J j satisfy J i < J j .We also require that each such x defines a sequence <strong>of</strong> natural numbers{n i } l+1i=1l+1 ∑such that n i = n, <strong>and</strong> the projections <strong>of</strong> the points x 1 , . . . , x n1 on Ti=1lie on the edge e 1 , the projections <strong>of</strong> the points x n1 +1, . . . , x n1 +n 2on T lie onthe edge e 2 <strong>and</strong> so on; <strong>and</strong> the projections <strong>of</strong> the points x n1 +...+n l +1, . . . , x n onT equal <strong>to</strong> the root <strong>of</strong> T . We denote the set <strong>of</strong> such n-tuples <strong>of</strong> points withsuch an induced sequence {n i } l+1i=1 , by X T,n 1 ,...,n l+1. We denoteX T,n := ∪ X T,n1 ,...,n l+1,where the union is over the set <strong>of</strong> l + 1-tuples <strong>of</strong> numbers n 1 , . . . , n l+1 such thatl+1 ∑n i = n.i=1Let h t be a Hamil<strong>to</strong>nian flow generated by H. Take p ∈ N. Then for eachx ∈ X T,nγ(h p 1; x) = α 1,p γ 1,p (h 1 ; x) · . . . · γ l,p (h 1 ; x)α 2,p , (3.9)for some γ i,p (h 1 ; x) ∈ P n , which pairwise commute with each other, <strong>and</strong> for somebraids α 1,p <strong>and</strong> α 2,p . The braids α 1,p ∈ B n <strong>and</strong> α 2,p ∈ B n depend on x ∈ X T,n ,p ∈ N <strong>and</strong> H. For all n ∈ N there exists a constant M(n) > 0, which dependsonly on n, such that for each x ∈ X T,n as above <strong>and</strong> for each p ∈ N this constantsatisfies the following inequations: |˜φ(α 1,p )| ≤ M(n), |˜φ(α 2,p )| ≤ M(n). Eachγ i,p (h 1 ; x) depends on the integer p, the function H, the pointsx n1 +n 2 +...+n i−1 +1, . . . , x n1 +n 2 +...+n i,<strong>and</strong> on the position <strong>of</strong> the points whose projections on T do not lie on the edgee i .Let η i,k,n be a pure braid in P n , in which the (i + k)-th str<strong>and</strong> makes onetwist in the positive direction around previous i − 1 str<strong>and</strong>s, <strong>and</strong> let η − i,k,n be a44


kiFigure 3.2: The braid η i,k,nn−k−ipure braid in P n , in which the (i + k)-th str<strong>and</strong> makes one twist in the negativedirection around previous i − 1 str<strong>and</strong>s, see Figure 3.2.If k = 0, then, by definition, η i,0,n = η i,n . We setAP l,n = {η i,k,n , η − i,k,n , | 2 ≤ i + k ≤ n}.Then AP l,n generates a subgroup <strong>of</strong> P n . Note that each braid η i,k,n ∈ AP l,n isconjugate in B n <strong>to</strong> the braid η i,n . Thus ˜φ(η i,k,n ) = ˜φ(η i,n ) (here we use the factthat ˜φ is B n -invariant). Note that each γ i,p (h 1 ; x) is equal <strong>to</strong> η j i,k,nfor some j,where i, j, k, n depend on p, x <strong>and</strong> H.For β ∈ P k we define β k,m ∈ P k+m <strong>to</strong> be the image <strong>of</strong> β under the st<strong>and</strong>ardinclusion <strong>of</strong> P k in<strong>to</strong> P m+k . It means that β k,m = β ⊔ Id m , where the identitybraid Id m ∈ P m is placed on the right <strong>of</strong> the braid β ∈ P k . Denoteβ(i, p) := γ i,p (h 1 ; x).Note that each γ i,p (h 1 ; x) is conjugate in B n <strong>to</strong> the braid β(i, p) ni ,n−n i. Thusthe conjugacy class in B n <strong>of</strong> each γ i,p (h 1 ; x) depends only on the integer p, thefunction H <strong>and</strong> the points x n1 +n 2 +...+n i−1 +1, . . . , x n1 +n 2 +...+n i.Note that any homogeneous <strong>quasi</strong>-morphism restricts <strong>to</strong> a homomorphismon an abelian subgroup. Recall that ˜φ is a B n -invariant <strong>quasi</strong>-morphism. Itfollows that for each x ∈ X T,n1limp→∞ p ˜φ(γ(hp 1; x)) =∑l+1i=1limp→∞1p ˜φ(γ i,p(h 1 ; x)) (3.10)45


<strong>and</strong>∫˜Φ(h 1 ) =X n∑l+1n!1∑l+1limp→∞ p ˜φ(γ(hp 1; x))dx = n!n∑n−n∑ 1i=1 n 1 =0 n 2 =0n−n∑1 −...−n l· · ·n l+1 =0∫i=1X T,n∫X T,n1 ,...,n l+11limp→∞ p ˜φ(γ i,p(h 1 ; x))dx =1limp→∞ p ˜φ(β(i, p) n i ,n−n i)dx. (3.11)We also see that for each fixed n in!n∑n−n∑ 1n 1 =0 n 2 =0n−n∑1 −...−n l· · ·n l+1 =0∫X T,n1 ,...,n l+1( nn i)vol(X Ai ,n i)1limp→∞ p ˜φ(β(i, p) n i ,n−n i)dx =∫X Ai ,n i1limp→∞ p ˜φ(β(i, p) n i ,n−n i)dx ni , (3.12)where dx ni = dx n1 +n 2 +...+n i−1 +1 · . . . · dx n1 +n 2 +...+n i−1 +n i, A i is an annulus whichcorresponds <strong>to</strong> an edge e i , X Ai ,n i(respectively X Ai ,n i) is the space <strong>of</strong> ordered n i -tuples <strong>of</strong> different points in A i (respectively (n − n i ) tuples <strong>of</strong> points in D 2 \ A i )such that this n i -tuple (respectively (n−n i )-tuple) is a part <strong>of</strong> x ∈ X T,n . Denotethe integral in the right-h<strong>and</strong> side in (3.12) by I i . Thus equations (3.11) <strong>and</strong>(3.12) tell us that in order <strong>to</strong> compute ˜Φ(h 1 ) it is enough <strong>to</strong> know the value I ifor each 1 ≤ i ≤ l + 1. We will compute I i by using angle-action coordinateson the domain A i .Step 2. We prove this theorem by induction on the number l <strong>of</strong> edges <strong>of</strong>the corresponding Reeb graph. If l = 1, then the pro<strong>of</strong> is exactly the same asin [26, Lemma 5.2]. The induction hypothesis states that our theorem is truefor every Reeb graph T with less than l edges.Let T be the Reeb graph <strong>of</strong> H which has l edges, <strong>and</strong> denote by e l theedge which connects the root <strong>of</strong> T <strong>to</strong> other vertex v l . Let k H <strong>and</strong> l − k H − 1be the number <strong>of</strong> edges in two connected components T 1 <strong>and</strong> T 2 <strong>of</strong> the graphT \ {e l , root, v l }. We cut the disc in<strong>to</strong> l + 1 pieces, by cutting it along it criticallevel curves, which passes through the critical points, <strong>and</strong> by cutting it along acurve c. We get l +1 open annuli. Denote the k H open annuli, corresponding <strong>to</strong>edges <strong>of</strong> T 1 , by A 1 , . . . , A kH , <strong>and</strong> denote the l−k H −1 open annuli, corresponding<strong>to</strong> edges <strong>of</strong> T 2 , by A kH +1, . . . , A l−1 . The open annulus A l corresponds <strong>to</strong> an edge46


A lA l +1curve cFigure 3.3: Curve c <strong>to</strong>gether with annuli A l <strong>and</strong> A l+1e l , i.e. the curve c is a connected component <strong>of</strong> ∂A l . We denote by A l+1 theother open annulus adjacent <strong>to</strong> c, see Figure 3.3.Let (J i , θ i ) be angle-action coordinates on A i for 1 ≤ i ≤ l, <strong>and</strong> let (J l+1 , θ l+1 )be the coordinates on A l+1 , where J l+1 is the area coordinate divided by 2π <strong>and</strong>θ l+1 is the angle coordinate. Denote by ε the area <strong>of</strong> A 0 divided by 2π. Nowwe can rewrite (3.11) as follows:∑l+1˜Φ(h 1 ) = n!(2π) nwhere Seq = {n i } l+1i=1n∑n−n∑ 1i=1 n 1 =0 n 2 =0· · ·n−n∑1 −...−n ln l+1 =0∫∫· · ·D Seq,l1limp→∞ p ˜φ(β(i, p) n i ,n−n i)dJ,l+1 ∑is a sequence <strong>of</strong> natural numbers such thati=1(3.13)n i = n <strong>and</strong>D Seq,l ⊂ [ 0, 1 2] ×nis the domain which is determined by the following inequalities:CR i < J i,1 < . . . < J i,ni < CR i + R i for each 1 ≤ i ≤ l − 1,R 1 + . . . + R l−1 < J l,1 < . . . < J l,nl < 1 2 − ε,12 − ε < J l+1,1 < . . . < J l+1,nl+1 < 1 2 ,47


<strong>and</strong> dJ = dJ 1,1 · . . . · dJ 1,n1 · . . . · dJ l+1,nl +1 · . . . · dJ l+1,n . We have <strong>to</strong> multiply by(2π) n in the formula above, because there are n angle coordinates.Step 3. Let H 1 = H| A1 ∪...∪A kH<strong>and</strong> H 2 = H| AkH +1∪...∪A l−1be Morse-typefunctions with Reeb graphs T 1 <strong>and</strong> T 2 respectively. We set(kH)(∑l−1) ∑R 1,l := R i , R 2,l := R i , R := R 1,l + R 2,l .i=1i=k H +1Denote( ) n − k˜φ i,k (J j ) := ˜φ(η i,n )i (J j ) i−1 (R 1,l − J j )in−k−i ∂∂J j(J j ),Ψ n (A j ) := (2π) nn∑( ) ( CR k ∫ j +R jn−kn 11,l)k 2 − R ∑˜φ i,k (J j )dJ jk=0CR jfor 1 ≤ j ≤ k H , <strong>and</strong> denote( ) n − k˜φ i,k (J j ) := ˜φ(η i,n )i (J j ) i−1 (R 2,l − J j )iΨ n (A j ) := (2π) ni=0n−k−i ∂∂J j(J j ),n∑( ) ( CR k ∫ j +R jn−kn 12,l)k 2 − R ∑˜φ i,k (J j )dJ jk=0CR jfor k H + 1 ≤ j ≤ l − 1. We denote( ) n − d˜φ i,k,d (J l ) = ˜φ(η i,n ) i(i − 1) · . . .i( ) n−d−i 1. . . · (i + k + d − n)(J l − R) i+k+d−n−1 2 − J ∂l − ε (J l ),∂J lwhere i + k + d − n ≥ 1 <strong>and</strong> setΨ n (A l ) := (2π) nn∑d=0( nd)ε d n−d∑k=1R n−d−k(n − d − k)!12 −ε ∫Ri=0∑n−di=n−d−k+1˜φ i,k,d (J l )dJ l .48


It follows from the induction hypothesis <strong>and</strong> (3.13) that˜Φ n (h 1 ) =l∑Ψ n (A j ).j=1An easy computation shows thatCR∫j +R jΨ n (A j ) = (2π) nfor 1 ≤ j ≤ l − 1, <strong>and</strong>CR j1∫2 −εΨ n (A l ) = (2π) nRn∑i=2n∑i=2( ) ( ) n−i n 1˜φ(η i,n )i (J j ) i−1i 2 − J ∂j (J j )dJ j∂J j( ) ( ) n−i n 1˜φ(η i,n )i (J l ) i−1i 2 − J ∂l (J l )dJ l .∂J lIt follows that∫˜Φ(h 1 ) =T ′ dµ.3.3 Properties <strong>of</strong> induced <strong>quasi</strong>-morphisms onthe group DIn this section we prove some properties the <strong>quasi</strong>-morphisms on D induced by<strong>quasi</strong>-morphisms on P n .Proposition 3.3.1. For each n ≥ 2 the homogeneous <strong>quasi</strong>-morphism ˜Φ n ,defined by a homogeneous <strong>quasi</strong>-morphism ˜φ : P n → R, is not continuous in theC 0 -<strong>to</strong>pology on D.Pro<strong>of</strong>. The case n = 2 was proved by Gambaudo <strong>and</strong> Ghys in [25]. Let n > 2.For each k ∈ N, k > 2, we define H k : D 2 → R as follows:1. H k (p, q) = h k (J), where J = p2 +q 2.22. h k (J) = 0 for J > 1. For each k there exist 0 < ε k k < 1 such that h k k(J) = C k49


for J ∈ [0, ε k ], where C k ≠ 0. The function h k (J) has no critical points in (ε k , 1 k ).3. The integral1∫20(2π) nn∑i=2( ) ( ) n−i n 1˜φ n (η i,n )i (J) i−1i 2 − J h ′ k(J)dJ = N k ,where N k ≥ 1 for all k > 2. We denote by h k,1 the time one Hamil<strong>to</strong>nian flow<strong>of</strong> H k . It is obvious that h k,1 converges <strong>to</strong> Id in C 0 -<strong>to</strong>pology. It follows fromthe pro<strong>of</strong> <strong>of</strong> Theorem 3.2.4 that ˜Φ n (h k,1 ) = N k ≥ 1. But ˜Φ n (Id) = 0.The goal <strong>of</strong> the remaining part <strong>of</strong> this section is <strong>to</strong> show that if H k −−−→ Hk→∞in C 1 -<strong>to</strong>pology, thenlim ˜Φ n (h 1,k ) = ˜Φ n (h 1 ).k→∞Here H k , H ∈ H 0 , <strong>and</strong> h 1,k , h 1 are Hamil<strong>to</strong>nian time-one flows generated by H k<strong>and</strong> H respectively.Let g ∈ D <strong>and</strong> ψ t : [0, 1] → D be such that ψ 0 = Id <strong>and</strong> ψ 1 = g. We denoteG({ψ t }) :=∫D 2 ×D 2 12π∫ 10∂∥∂t( )∥ψt (x) − ψ t (y) ∥∥∥dtdxdy.∥ψ t (x) − ψ t (y)∥Lemma 1 in [28] shows that G({ψ t }) is well defined. DenoteG(g) := inf G({ψ t }),where the infimum is taken over all iso<strong>to</strong>pies in D joining the identity with g.Lemma 2 in [28] states that for all g <strong>and</strong> h in DThus the following limit exists:G(gh) ≤ G(g) + G(h).G(g i )L(g) = lim .i→∞ iProposition 3.3.2. Let g ∈ D, <strong>and</strong> let ˜φ : P n → R be a homogeneous <strong>quasi</strong>morphism.Then ∣ ∣∣˜Φn (g)∣∣ ≤ K 1 L(g),where K 1 > 0 is independent <strong>of</strong> g.50


Pro<strong>of</strong>. Apply equation (3.7) in Lemma 3.1.1 <strong>to</strong> g p instead <strong>of</strong> h 1 <strong>and</strong> take theinfimum <strong>of</strong> the expression in the right-h<strong>and</strong> side over all the paths connectingthe identity with g p . The resulting infimum has the form K 1 G(g p ) + K 2 , whereK 1 <strong>and</strong> K 2 are positive constants independent <strong>of</strong> g. Thus|Φ(g p )| ≤ K 1 G(g p ) + K 2 ,for any p. Dividing both sides by p <strong>and</strong> taking the limit as p → +∞ we get therequired inequality.Now we will define a metric on D. A tangent vec<strong>to</strong>r <strong>to</strong> D at a point g is aHamil<strong>to</strong>nian vec<strong>to</strong>r field Y g (see [22]) whose L 2 -norm is defined by⎛∫∥Y g ∥ 2 := ⎝D 2⎞∥Y g (x)∥ 2 dx⎠where ∥.∥ st<strong>and</strong>s for the st<strong>and</strong>ard Euclidean norm. For any path ψ t in D whichconnects g <strong>and</strong> h, the length <strong>of</strong> ψ t is given by the formula:We denotel({ψ t }) =∫ 10dψ t∥ dt ∥ dt.2d 2 (g, h) := inf l({ψ t }),where the infimum is taken over all paths in D joining g with h. One can showthat d 2 is a distance function. Thus (D, d 2 ) is a metric space. Note that d isright invariant: d 2 (g, h) = d 2 (gf, hf).Corollary 3.3.3. Let g ∈ D, <strong>and</strong> let ˜φ : P n → R be a homogeneous <strong>quasi</strong>morphism.Then|˜Φ n (g)| ≤ K 3 d 2 (Id, g),where K 3 > 0 depends only on n.Pro<strong>of</strong>. It follows from [28, Theorem 1] that for any g ∈ D there exists a universalconstant K > 0, such thatL(g) ≤ Kd 2 (Id, g).Now take K 3 = K · K 1 . The statement follows from Proposition 3.3.2.5112,


Lemma 3.3.4. Let F ∈ H 0 . Then for any ε > 0 <strong>and</strong> p ∈ N there exists δ p > 0,such that if H ∈ H 0 is δ p -close <strong>to</strong> F in C 1 -<strong>to</strong>pology, thend 2 (f p 1 , h p 1) < ε,where f t <strong>and</strong> h t are the Hamil<strong>to</strong>nian flows generated by F <strong>and</strong> H.Pro<strong>of</strong>. For the convenience we normalize the area <strong>of</strong> D 2 <strong>to</strong> be 1. First, we willshow that for all p ∈ N there exists δ 1,p > 0 such thatNote thatmax ∥∇Fx∈D 2 (x) − ∇H (x) ∥ < δ 1,p ⇒ d 2 (f p 1 , h p 1) < ε.d 2 (f p 1 , h p 1) = d 2 (Id, f p 1 h −p1 ) ≤ l({f p t h −pt }).It follows from [51, Proposition 1.4.D] that∂(f p t h −p∂tt )(x) = p · (sgrad(F ) − sgrad(Hf −pt )) (fpt h−p t (x)) ,where sgrad is the symplectic gradient. Thus∂(f p t h −pt )∥∥ (x)∥∥(sgrad(F∂t ∥ = p · ) − sgrad(Hf−pt∥= p · ∥(∇F − ∇(Hf −p)) (fp ∥ ,tt h−p t (x)))) (fpt h−p t (x))Note that f t is an au<strong>to</strong>nomous Hamil<strong>to</strong>nian flow. Thus F f t (x) = F (x) for allx ∈ D 2 <strong>and</strong> t ∈ R. It follows that ∀x ∈ D 2 <strong>and</strong> ∀p ∈ Z∇F (h−pt (x))(Df−pWe get the following inequality:∫ 1d 2 (f p 1 , h p 1) ≤ p0⎛∫⎝D 2t∥(Df −pt ) (fp) (fpt h−p tt h−p t(x)) = ∇F (f p t h−p t (x)) .(x)) ∥2 ∥∇F (h−pt∥⎞ 12(x)) − ∇H (h −pt (x)) ∥2 ⎠ dt.DenoteM Dft :=max ∥(Dfx∈D 2 t −1 ) (x) ∥.,t∈[0,1]52


TakeWe get the following inequalityδ 1,p =εp(M Dft ) p .d 2 (f p 1 , h p 1) ≤ p(M Dft ) p max ∥∇F (x) − ∇H(x)∥ ≤ ε.x∈D2 Denote by δ p = δ 1,p2 . We have shown that if H is δ p-close <strong>to</strong> F in C 1 -<strong>to</strong>pology,then d 2 (f p 1 , h p 1) < ε.Theorem 3.3.5. Let F ∈ H 0 . Then for any ε > 0 there exists δ > 0, such thatif H ∈ H 0 is δ-close <strong>to</strong> F in C 1 -<strong>to</strong>pology then:∣∣˜Φ n (f 1 ) − ˜Φ n (h 1 ) ∣ ≤ ε,where f 1 <strong>and</strong> h 1 are the time-one Hamil<strong>to</strong>nian flows generated by F <strong>and</strong> H.Pro<strong>of</strong>. Let ε > 0, let D˜Φnbe the defect <strong>of</strong> the homogeneous <strong>quasi</strong>-morphism˜Φ n : D → R,<strong>and</strong> let K 3 be the constant which was defined in Corollary 3.3.3. Take p ∈ Nsuch that D˜Φn +K 3< ε. It follows from Lemma 3.3.4 that there exists δp p > 0,such that if H is δ p -close <strong>to</strong> F in C 1 -<strong>to</strong>pology, then d 2 (f p 1 , h p 1) < 1. Thus we getthe following inequality:∣∣˜Φ n (f 1 ) − ˜Φ n (h 1 ) ∣ = 1 ∣∣˜Φ n (f p 1 ) −p˜Φ ∣n (h p + ∣˜Φ D˜Φn n (f p 1 h −p1 ) ∣1) ∣ ≤.pIt follows from Corollary 3.3.3 that∣∣˜Φ n (f p 1 h −p1 ) ∣ ≤ K 3 d 2 (Id, f p 1 h −p1 ) = K 3 d 2 (f p 1 , h p 1) < K 3 .Thus ∣ ∣∣˜Φn (f 1 ) − ˜Φ n (h 1 )∣∣ < D˜Φn + K 3p< ε.53


3.4 Induced <strong>quasi</strong>-morphisms <strong>and</strong> Calabihomomorphism3.4.1 Rasmussen <strong>quasi</strong>-morphisms on DLet ˜s be the homogeneous <strong>quasi</strong>-morphism on P n induced by the Rasmussenknot invariant s. The induced family <strong>of</strong> homogeneous <strong>quasi</strong>-morphisms on D iswell defined for every integer n ≥ 2, this family is denoted by ˜Ras n .Lemma 3.4.1. Let x, a ∈ R. Then for every n ∈ Nn(n − 1)x(x + a) n−2 =n∑i=2( ) ni(i − 1) x i−1 a n−i . (3.14)iPro<strong>of</strong>. Denote by f(x) = (x + a) n <strong>and</strong> by f ′ the derivative <strong>of</strong> f. The pro<strong>of</strong>follows from the fact that n(n − 1)x(x + a) n−2 = xf ′′ (x) <strong>and</strong> that(x + a) n =n∑i=0( ni)x i a n−i .Theorem 3.4.2. For every g ∈ D˜Ras n (g) = 2n(n − 1)π n−1 · C(g). (3.15)Pro<strong>of</strong>. The theorem <strong>of</strong> Banyaga [5] states that ker(C) is a simple group, whereC : D → R is a Calabi homomorphism. Thus there exists a homomorphismκ : R → R such that for all g ∈ Dκ(C(g)) = ˜Ras n (g). (3.16)Restricting ˜Ras n <strong>and</strong> C on a 1-parametric subgroup <strong>of</strong> D, on which they do notvanish, one easily sees that κ is continuous. It means thatM n · C(g) = ˜Ras n (g), (3.17)where M n is some real constant which depends only on n.54


Take H : D 2 → R, such that H equals zero near the boundary, <strong>and</strong> suchthat H has one maximum point <strong>and</strong> has no other critical points. Denote by h 1the time-one Hamil<strong>to</strong>nian flow generated by H. Recall that∫C(h 1 ) = −2 H.D 2ThusC(h 1 ) = −4π1∫20(J)dJ,where J is the symplectic action coordinate. It follows from Corollary 2.2.3that˜s(η i,n ) = lk(η i,n ) = 2(i − 1).It follows from Lemma 3.4.1 thatn∑( ) ( ) n−i n 1(2π) n ˜s(η i,n )i J i−1i 2 − J = 8π n n(n − 1)J.i=2It follows from Theorem 3.2.4 thatNote that ( 1 ) = 0. Thus2˜Ras n (h 1 ) = 8π n n(n − 1)˜Ras n (h 1 ) = −8π n n(n − 1)It follows that M n = 2n(n − 1)π n−1 .1∫201∫20J ′ (J)dJ.(J)dJ = 2n(n − 1)π n−1 C(h 1 ).3.4.2 Signature <strong>quasi</strong>-morphisms <strong>and</strong> CalabihomomorphismRecall that ˜sign n : B n → R is a homogeneous <strong>quasi</strong>-morphism for all n ≥ 2.The induced family <strong>of</strong> homogeneous <strong>quasi</strong>-morphisms is denoted by ˜Sign n,D 2.55


In [26] Gambaudo <strong>and</strong> Ghys proved that all <strong>of</strong> them are non-trivial <strong>and</strong> linearlyindependent. In the case where the Reeb graph <strong>of</strong> H is simple they showed that∫˜Sign n,D 2(h 1 ) = ′ dµ.Proposition 3.4.3. For each natural number n we have∫(˜Sign n,D 2(h 1 ) = nπ ) n 1 + 4(n − 1)J − (1 − 4J)n−1 ′ (J)dJ, (3.18)Twhere h 1 is the time-one Hamil<strong>to</strong>nian flow defined by a Morse-type Hamil<strong>to</strong>nianH.First we prove the following combina<strong>to</strong>rial equality.Lemma 3.4.4. Let x, a ∈ R. Then for every n ∈ Nn∑i=2n2( ) ( ) n−i n 1˜sign n (η i,n )i x i−1i 2 − x =((x + a) n−1 + 2(n − 1)x(x + a) n−2 − (a − x) n−1) .Pro<strong>of</strong>. Recall that in Lemma 2.2.10 we proved that{i, if i is even,˜sign n (η i,n ) =i − 1, if i is odd.It follows thatn∑( ) n˜sign n (η i,n )i x i−1 a n−i =ii=2n∑i=2( ) ni(i − 1) x i−1 a n−i + 1 i 2n(n − 1)x(x + a) n−2 + n 2n∑i=1T( ) n (1i) − (−1)i−1x i−1 a n−i =i((x + a) n−1 − (a − x) n−1) .56


Pro<strong>of</strong> <strong>of</strong> Proposition 3.4.3. We change variables x = J <strong>and</strong> a = 1 −J in Lemma23.4.4. It follows thatn∑( ) ( ) n−i n 1˜sign n (η i,n )i J i−1i 2 − J =i=2( (1 ) n−1 ( ) n−1 ( ) ) n−1n1 1+ 4(n − 1)J −2 22 2 − 2J . (3.19)Now the pro<strong>of</strong> follows easily from Theorem 3.2.4 <strong>and</strong> (3.19).We are finally ready <strong>to</strong> prove the second claim in Proposition 2.2.9.Pro<strong>of</strong> <strong>of</strong> Proposition 2.2.9. Claim 2. Suppose that ˜sign n : P n → R is a homomorphismfor n > 2. Then ˜Sign n,D 2 : D → R is a homomorphism. LetC : D → R be the Calabi homomorphism. Using the theorem <strong>of</strong> Banyaga [5],which states that ker(C) is a simple group, we get that there exists a homomorphismκ : R → R such that˜Sign n,D 2 = κC.Restricting ˜Sign n,D 2 <strong>and</strong> C on a 1-parametric subgroup <strong>of</strong> D, on which they donot vanish, one easily sees that κ is continuous. Hence κ(y) = K 1,n y for ally ∈ R, <strong>and</strong>˜Sign n,D 2 = K 1,n C,where K 1,n is some real constant which depends only on n. But using Proposition3.4.3, we see that this is impossible for n > 2.Theorem 3.4.5. For each g ∈ D generated by an au<strong>to</strong>nomous Hamil<strong>to</strong>nianlimn→∞˜Sign n,D 2(g)π n−1 n(n − 1) = C(g).Pro<strong>of</strong>. Step 1. Suppose that g is generated by a Morse-type Hamil<strong>to</strong>nianH ∈ H 0 . It follows from (3.18) thatlimn→∞˜Sign n,D 2(g)π n−1 n(n − 1) = C(g).Step 2. Suppose that g is generated by any H ∈ H 0 . Then there exists aMorse-type function F ∈ H 0 , which is C 1 -close <strong>to</strong> H (see [46, Theorem 2.7]).Now the statement follows from Theorem 3.3.5.57


3.5 Open problemsThe following three open problems arise naturally in connection with our results.Problem 1. Does there exist a real-valued knot invariant I, independentfrom the ω-signatures, τ <strong>and</strong> the Rasmussen invariant s, such that I satisfies theconditions <strong>of</strong> Theorem 2.1.4, <strong>and</strong> such that the homogenization <strong>of</strong> the induced<strong>quasi</strong>-morphism Î : P n → R is non-trivial?Problem 2. Surprisingly, the pro<strong>of</strong> <strong>of</strong> the second claim in Proposition 2.2.9– dealing with a <strong>quasi</strong>-morphism on P n – involves the Calabi homomorphismon D. It would be interesting <strong>to</strong> find a direct pro<strong>of</strong> <strong>of</strong> this claim.Problem 3. The following problem is motivated by a question posed <strong>to</strong>us by L.Polterovich. Denote by D aut ⊂ D the set <strong>of</strong> area-preserving diffeomorphismsgenerated by au<strong>to</strong>nomous Hamil<strong>to</strong>nians.Observe that, by our results in subsection about signature <strong>quasi</strong>-morphisms,one can find a linear combination Φ <strong>of</strong> homogeneous <strong>quasi</strong>-morphisms ˜sign ω onB n so that the corresponding homogeneous <strong>quasi</strong>-morphism ˜Φ : D → R vanisheson D aut .On the other h<strong>and</strong>, using a construction from [26] (completely different fromthe one described in this paper) one can construct more homogeneous <strong>quasi</strong>morphisms˜Ψ α on D that vanish on D aut .It would be interesting <strong>to</strong> check whether the <strong>quasi</strong>-morphism ˜Φ is non-trivial<strong>and</strong>, if so, whether it is a linear combination <strong>of</strong> the <strong>quasi</strong>-morphisms ˜Ψ α .So far we have not been able <strong>to</strong> compute explicitly the value <strong>of</strong> any <strong>of</strong> theGambaudo-Ghys <strong>quasi</strong>-morphisms coming from <strong>quasi</strong>-morphisms on B n on anyarea-preserving diffeomorphism that is not generated by a Hamil<strong>to</strong>nian flowpreserving a foliation.58


Chapter 4Link <strong>invariants</strong> via countingsurfacesA Gauss diagram is a simple, combina<strong>to</strong>rial way <strong>to</strong> present a knot. It is knownthat any Vassiliev invariant may be obtained from a Gauss diagram formulathat involves counting (with signs <strong>and</strong> multiplicities) subdiagrams <strong>of</strong> certaincombina<strong>to</strong>rial types. These formulas generalize the calculation <strong>of</strong> a linkingnumber by counting signs <strong>of</strong> crossings in a link diagram. Until recently, explicitformulas <strong>of</strong> this type were known only for few <strong>invariants</strong> <strong>of</strong> low degrees.In this chapter we present simple formulas for two infinite families <strong>of</strong> <strong>invariants</strong>in terms <strong>of</strong> counting surfaces <strong>of</strong> a certain genus <strong>and</strong> number <strong>of</strong> boundarycomponents in a Gauss diagram. The first family <strong>of</strong> <strong>invariants</strong> is a Chmu<strong>to</strong>v-CapKhoury-Rossi family generalized <strong>to</strong> classical <strong>and</strong> long virtual links witharbitrary number <strong>of</strong> components. The resulting <strong>invariants</strong> are identified withcoefficients <strong>of</strong> the Conway polynomial. The second family <strong>of</strong> <strong>invariants</strong> is identifiedwith certain derivatives <strong>of</strong> the HOMFLYPT polynomial. In the remainingpart <strong>of</strong> the thesis we consider only oriented links.4.1 Gauss diagrams <strong>and</strong> arrow diagramsIn this section we recall a notion <strong>of</strong> Gauss diagrams, arrow diagrams <strong>and</strong> Gaussdiagram formulas. We then define a special type <strong>of</strong> arrow diagrams whichwill be used <strong>to</strong> define Gauss diagram formulas for coefficients <strong>of</strong> the Conwaypolynomial, <strong>and</strong> for coefficients <strong>of</strong> some other polynomial derived from theHOMFLYPT polynomial.59


4.1.1 Gauss diagrams <strong>of</strong> classical <strong>and</strong> virtual linksGauss diagrams (see e.g. [30], [53]) provide a simple combina<strong>to</strong>rial way <strong>to</strong>encode classical <strong>and</strong> virtual links.Definition 4.1.1. Given a classical (possibly framed) link diagram D, considera collection <strong>of</strong> oriented circles parameterizing it. Unite two preimages <strong>of</strong> everycrossing <strong>of</strong> D in a pair <strong>and</strong> connect them by an arrow, pointing from the overpassingpreimage <strong>to</strong> the underpassing one. To each arrow we assign a sign (localwrithe) <strong>of</strong> the corresponding crossing. The result is called the Gauss diagramG corresponding <strong>to</strong> D.We consider Gauss diagrams up <strong>to</strong> an orientation-preserving diffeomorphisms<strong>of</strong> the circles. In figures we will always draw circles <strong>of</strong> the Gauss diagramwith a counter-clockwise orientation.Example 4.1.2. Diagrams <strong>of</strong> the trefoil knot <strong>and</strong> the Hopf link, <strong>to</strong>gether withthe corresponding Gauss diagrams, are shown in the following picture.+++A classical link can be uniquely reconstructed from the corresponding Gaussdiagram [30]. Many fundamental knot <strong>invariants</strong>, such as the knot group <strong>and</strong>the Alex<strong>and</strong>er polynomial, may be easily obtained directly from the Gaussdiagram. We are going <strong>to</strong> work with based Gauss diagrams, i.e. Gauss diagramswith a base point (different from the endpoints <strong>of</strong> the arrows) on one <strong>of</strong> thecircles. If we cut a based circle at the base point, we will get a Gauss diagram<strong>of</strong> a long link, see Figure 4.1.++++++Figure 4.1: Diagrams <strong>of</strong> based <strong>and</strong> long classical Hopf links <strong>to</strong>gether with theassociated Gauss diagrams60


+++Figure 4.2: Virtual trefoil <strong>and</strong> a virtual Hopf link with the corresponding GaussdiagramsNote that not every collection <strong>of</strong> circles with signed arrows is realizable asa Gauss diagram <strong>of</strong> a classical link, see Figure 4.2.The Gauss diagram <strong>of</strong> a virtual link diagram is constructed in the sameway as for a classical link diagram, but all virtual crossings are disregarded,see Figures 4.2 <strong>and</strong> 4.3. Each non-realizable Gauss diagram with a base pointrepresents a long virtual link, see Figure 4.3.++++Figure 4.3: Based <strong>and</strong> long virtual trefoil with the corresponding Gauss diagramsTwo Gauss diagrams represent iso<strong>to</strong>pic classical/virtual links (long links)if <strong>and</strong> only if they are related by a finite number <strong>of</strong> Reidemeister moves forGauss diagrams (applied away from the base point) shown in Figure 4.4, whereε = ±1. See e.g. [16, 48, 52].Two Gauss diagrams represent iso<strong>to</strong>pic classical framed links if <strong>and</strong> only ifthey are related by a finite number <strong>of</strong> Reidemeister moves for framed Gaussdiagrams. It suffices <strong>to</strong> consider Ω 2 <strong>and</strong> Ω 3 <strong>of</strong> Figure 4.4 <strong>and</strong> substitute the61


Ω 1 : ε Ω 2 :ε−εΩ 3 :++++++Figure 4.4: Reidemeister moves <strong>of</strong> Gauss diagramsmove Ω 1 byεεΩ F 1 : −ε−εNote that segments involved in Ω 2 or Ω 3 may lie on different components <strong>of</strong> thelink <strong>and</strong> the order in which they are traced along the link may be arbitrary.4.1.2 Arrow diagrams <strong>and</strong> Gauss diagram formulasAn arrow diagram is a modification <strong>of</strong> a notion <strong>of</strong> a Gauss diagram, in whichwe forget about realizability <strong>and</strong> signs <strong>of</strong> arrows, see Figure 4.5. In other words,Figure 4.5: Connected arrow diagramsan arrow diagram consists <strong>of</strong> a number <strong>of</strong> oriented circles with several arrowsconnecting pairs <strong>of</strong> distinct points on them. We consider these diagrams up<strong>to</strong> orientation-preserving diffeomorphisms <strong>of</strong> the circles. An arrow diagram isbased, if a base point (different from the end points <strong>of</strong> the arrows) is markedon one <strong>of</strong> the circles. An arrow diagram is connected, if it is connected as62


a graph. Further we will consider only based connected arrow diagrams, sowe will omit mentioning these requirements throughout this chapter, unless amisunderst<strong>and</strong>ing is likely <strong>to</strong> occur. In figures we will always draw the circles<strong>of</strong> an arrow diagram with a counter-clockwise orientation.M. Polyak <strong>and</strong> O. Viro suggested [53] the following approach <strong>to</strong> computelink <strong>invariants</strong> using Gauss diagrams.Definition 4.1.3. Let A be an arrow diagram with m circles <strong>and</strong> let G be abased Gauss diagram <strong>of</strong> an m-component oriented (long, virtual) link. A homomorphismϕ : A → G is an orientation preserving homeomorphism betweeneach circle <strong>of</strong> A <strong>and</strong> each circle <strong>of</strong> G, which maps a base point <strong>of</strong> A <strong>to</strong> the basepoint <strong>of</strong> G <strong>and</strong> induces an injective map <strong>of</strong> arrows <strong>of</strong> A <strong>to</strong> the arrows <strong>of</strong> G.The set <strong>of</strong> arrows in Im(ϕ) is called a state <strong>of</strong> G induced by ϕ <strong>and</strong> is denotedby S(ϕ). The sign <strong>of</strong> ϕ is defined as sign(ϕ) = ∏ α∈S(ϕ)sign(α). A set <strong>of</strong> allhomomorphisms ϕ : A → G is denoted by Hom(A, G).Note that since the circles <strong>of</strong> A are mapped <strong>to</strong> circles <strong>of</strong> G, a state S <strong>of</strong> Gdetermines both the arrow diagram A <strong>and</strong> the map ϕ : A → G with S = S(ϕ).Definition 4.1.4. A pairing between an arrow diagram A <strong>and</strong> G is defined by∑⟨A, G⟩ = sign(ϕ).ϕ∈Hom(A,G)For an arbitrary arrow diagram A the pairing ⟨A, G⟩ does not represent a linkinvariant, i.e. it depends on the choice <strong>of</strong> a Gauss diagram <strong>of</strong> a link. However,for some special linear combinations <strong>of</strong> arrow diagrams the result is independen<strong>to</strong>f the choice <strong>of</strong> G, i.e. does not change under the Reidemeister moves for Gaussdiagrams. Using a slightly modified definition <strong>of</strong> arrow diagrams Goussarov,Polyak <strong>and</strong> Viro showed in [30] that each real-valued Vassiliev invariant <strong>of</strong> longknots may be obtained this way. In particular, all coefficients <strong>of</strong> the Conwaypolynomial may be obtained using suitable combinations <strong>of</strong> arrow diagrams.4.1.3 Surfaces corresponding <strong>to</strong> arrow diagramsGiven an arrow diagram A, we define an oriented surface Σ(A) as follows.Firstly, replace each circle <strong>of</strong> A with an oriented disk bounding this circle.Secondly, glue 1-h<strong>and</strong>les <strong>to</strong> boundaries <strong>of</strong> these disks using each arrow as a core<strong>of</strong> a ribbon. See Figure 4.6.63


Figure 4.6: Constructing a surface from an arrow diagramDefinition 4.1.5. By the genus <strong>and</strong> the number <strong>of</strong> boundary components <strong>of</strong> anarrow diagram A we mean the genus <strong>and</strong> the number <strong>of</strong> boundary components<strong>of</strong> Σ(A).Remark 4.1.6. Let A be an arrow diagram with m circles <strong>and</strong> n arrows. Thenthe Euler characteristic χ <strong>of</strong> Σ(A) equals <strong>to</strong> χ(Σ(A)) = m−n. If A is connected,n ≥ m − 1. If A has one boundary component, n ≠ m(mod2).Example 4.1.7. The arrow diagram with one circle in Figure 4.5 is <strong>of</strong> genusone, while the other arrow diagram in the same figure is <strong>of</strong> genus zero. Both <strong>of</strong>them have one boundary component.Further we will work only with based connected arrow diagrams with oneor two boundary components.4.1.4 Ascending <strong>and</strong> descending arrow diagramsIn this subsection we define a special type <strong>of</strong> arrow diagrams with one <strong>and</strong> twoboundary components.Definition 4.1.8. Let A be a based arrow diagram with one boundary component.As we go along the boundary <strong>of</strong> Σ(A) starting from the base point,we pass on the boundary <strong>of</strong> each ribbon twice: once in the direction <strong>of</strong> itscore arrow, <strong>and</strong> once in the opposite direction. A is ascending (respectively,descending) if we pass each ribbon <strong>of</strong> Σ(A) first time in the direction opposite<strong>to</strong> its core arrow (respectively, in the direction <strong>of</strong> its core arrow).Remark 4.1.9. In order <strong>to</strong> define the notion <strong>of</strong> ascending (descending) arrowdiagrams we used the fact that all arrow diagrams are based <strong>and</strong> connected.The position <strong>of</strong> the base point in a connected arrow diagram is essential <strong>to</strong>define an order <strong>of</strong> the passage.Example 4.1.10. Arrow diagrams presented below are ascending (a), descending(b) <strong>and</strong> neither ascending nor descending (c).64


a b cDenote by A n,m (respectively, D n,m ) the set <strong>of</strong> all ascending (respectively,descending) arrow diagrams with n arrows, m circles <strong>and</strong> one boundary component.Example 4.1.11. The sets A 2,1 <strong>and</strong> D 2,1 are presented below.A 2,1 := <strong>and</strong> D 2,1 :=Definition 4.1.12. Let G be any Gauss diagram with m circles. We setA n,m (G) :=∑⟨A, G⟩ D n,m (G) := ∑⟨A, G⟩A∈A n,m A∈D n,m<strong>and</strong> define the following polynomials:∇ asc (G) :=∞∑A n,m (G)z n ∇ des (G) :=n=0∞∑D n,m (G)z nThese polynomials will play an important role in the next section. Now wegeneralize a notion <strong>of</strong> ascending (descending) arrow diagram <strong>to</strong> arrow diagramswith two boundary components.Definition 4.1.13. Let A be an arrow diagram with two boundary components.As we go along the component <strong>of</strong> ∂Σ(A) starting from the base point, we passon the boundary <strong>of</strong> each ribbon once or twice. We call core arrows, whichwe pass only in one direction, the separating arrows. Now we place anotherstarting point • on the second component <strong>of</strong> ∂Σ(A) near the first separatingarrow which we encounter in the passage, <strong>and</strong> start going along this componen<strong>to</strong>f ∂Σ(A). A is ascending (respectively, descending) if we pass each ribbon <strong>of</strong>Σ(A) first time in the direction opposite <strong>to</strong> its core arrow (respectively, in thedirection).n=065


Example 4.1.14. Arrow diagrams below have two boundary components. Diagram(a) is ascending <strong>and</strong> diagram (b) is descending. Separating arrows areshown in bold.aDenote by A 2 n,m (respectively, D 2 n,m) the set <strong>of</strong> all ascending (respectively,descending) arrow diagrams with n arrows, m circles <strong>and</strong> two boundary components.Example 4.1.15. All diagrams in the set A 2 2,2 are presented below.bLet G be any Gauss diagram with m circles. We setA 2 n,m(G) :=∑⟨A, G⟩ Dn,m(G) 2 := ∑⟨A, G⟩.A∈A 2 n,mA∈D 2 n,mA state S(ϕ) corresponding <strong>to</strong> ϕ : A → G for an ascending (respectivelydescending) diagram A with one or two boundary components will be also calledascending (respectively descending). It is useful <strong>to</strong> reformulate this notion interms <strong>of</strong> a tracing <strong>of</strong> G.Definition 4.1.16. Given ϕ : A → G, a passage along the boundary <strong>of</strong> thesurface Σ(A) induces a tracing <strong>of</strong> G: we follow an arc <strong>of</strong> a circle <strong>of</strong> G startingfrom the base point until we hit an arrow in S(ϕ), turn <strong>to</strong> this arrow, thencontinue on another arc <strong>of</strong> G following the orientation <strong>and</strong> so on, until wereturn <strong>to</strong> the base point. In case <strong>of</strong> two boundary components we repeat thesame procedure starting near the image <strong>of</strong> the first separating arrow. Then astate S(ϕ) is ascending (respectively descending), if we approach every arrowin the tracing first time at its head (respectively at its tail).4.1.5 Separating statesIn this subsection we define a notion <strong>of</strong> a separating state. This notion will beextensively used in the following sections.66


1 12 22 2 1 1iijja b cFigure 4.7: Ascending <strong>and</strong> descending labelingDefinition 4.1.17. Let G be a based Gauss diagram. An ascending (respectivelydescending) separating state S <strong>of</strong> G is a state S <strong>of</strong> G, <strong>to</strong>gether with alabeling <strong>of</strong> arcs <strong>of</strong> G (i.e., intervals <strong>of</strong> circles <strong>of</strong> G between endpoints <strong>of</strong> arrows)by 1 <strong>and</strong> 2 such that:1. Each arc near α ∈ S is labeled as in Figure 4.7a (respectively in Figure4.7b).2. Each arc near α /∈ S is labeled as in Figure 4.7c.3. An arc with a base point is labeled by 1.Every separating (ascending or descending) state S in G defines a new Gaussdiagram G S with labeled circles as follows:We smooth each arrow in G which belongs <strong>to</strong> S, as shown in Figure 4.8, <strong>and</strong>denote resulting smoothed Gauss diagram by G S . Each circle in G S is labeledby i, if it contains an arc labeled by i.Figure 4.8: Smoothing <strong>of</strong> an arrowNow we return <strong>to</strong> arrow diagrams with two boundary components. LetA ∈ A 2 n,m or A ∈ Dn,m. 2 We denote by σ(A) the set <strong>of</strong> separating arrows in A<strong>and</strong> label the arcs <strong>of</strong> circles in A by 1 if the corresponding arc belongs <strong>to</strong> thefirst boundary component <strong>of</strong> Σ(A) <strong>and</strong> by 2 otherwise.67


Note that each homomorphism ϕ : A → G induces an ascending or descendingseparating state S <strong>of</strong> G, by taking S = ϕ(σ(A)) <strong>and</strong> labeling each arc <strong>of</strong> Gby the same label as the corresponding arc <strong>of</strong> A.Definition 4.1.18. Let G be a based Gauss diagram with m circles. Let Sbe an ascending (respectively descending) separating state <strong>of</strong> G, A ∈ A 2 n,m(respectively A ∈ D 2 n,m), <strong>and</strong> ϕ : A → G. We say that ϕ is S-admissible, ifan ascending (respectively descending) separating state induced by ϕ coincideswith S.Definition 4.1.19. Let S be an ascending (respectively descending) separatingstate <strong>of</strong> G, <strong>and</strong> A ∈ A 2 n,m (respectively A ∈ Dn,m). 2 In each case we define anS-pairing ⟨A, G⟩ S by:⟨A, G⟩ S :=∑sign(ϕ),ϕ:A→Gwhere the summation is over all S-admissible ϕ : A → G. We setA 2 n,m(G) S :=∑⟨A, G⟩ S <strong>and</strong> Dn,m(G) 2 S := ∑⟨A, G⟩ S .A∈A 2 n,mA∈D 2 n,m4.2 Counting surfaces with one boundarycomponentIn this section we review Gauss diagram formulas for coefficients <strong>of</strong> the Conwaypolynomial ∇ obtained in [15, Theorem 3.5] for classical knots <strong>and</strong> 2-componentclassical links.Theorem 4.2.1 ([15]). Let G be a Gauss diagram <strong>of</strong> a classical knot or 2-component classical link L.where ∇(L) is the Conway polynomial.∇ asc (G) = ∇ des (G) = ∇(L), (4.1)Let G be a Gauss diagram with m circles. We give a direct pro<strong>of</strong> <strong>of</strong> theinvariance <strong>of</strong> both ∇ asc (G) <strong>and</strong> ∇ des (G) under the Reidemeister moves which donot involve a base point. This allows us <strong>to</strong> extend this result <strong>to</strong> m-componentclassical links <strong>and</strong> also <strong>to</strong> define two different generalizations <strong>of</strong> the Conwaypolynomial <strong>to</strong> long virtual links. At the end <strong>of</strong> this section we present someproperties <strong>of</strong> these polynomials.68


4.2.1 Invariants <strong>of</strong> long linksIn this subsection we generalize the result <strong>of</strong> [15] <strong>to</strong> m-component (classical orvirtual) long links.Theorem 4.2.2. Let G be a Gauss diagram <strong>of</strong> an m-component (classical orvirtual) long link L. Then ∇ asc (G) <strong>and</strong> ∇ des (G) define polynomial <strong>invariants</strong><strong>of</strong> long links, i.e. do not depend on the choice <strong>of</strong> G.Pro<strong>of</strong>. We will prove that ∇ asc (G) is an invariant <strong>of</strong> an underlying link. Thepro<strong>of</strong> for ∇ des (G) is the same. It suffices <strong>to</strong> show that A n,m (G) is invariant underthe Reidemeister moves Ω 1 , Ω 2 <strong>and</strong> Ω 3 <strong>of</strong> Figure 4.4 applied away from the basepoint. Let G <strong>and</strong> ˜G be two Gauss diagrams that differ by an application <strong>of</strong> Ω 1 ,so that ˜G has one additional isolated arrow α on one <strong>of</strong> the circles. Ascendingstates <strong>of</strong> G are in bijective correspondence with ascending states <strong>of</strong> ˜G which donot contain α. But α cannot not be in the image <strong>of</strong> ϕ : A → G with A ∈ A n,m ,because A should have one boundary component. ThusA n,m (G) = A n,m ( ˜G).Let G <strong>and</strong> ˜G be two Gauss diagrams that differ by an application <strong>of</strong> Ω 2 , sothat ˜G has two additional arrows α ε <strong>and</strong> α −ε , see Figure 4.4. Ascending states<strong>of</strong> G are in bijective correspondence with ascending states <strong>of</strong> ˜G which do notcontain α ±ε . Note that both α ε <strong>and</strong> α −ε can not be in the image <strong>of</strong> ϕ : A → Gwith A ∈ A n,m because A has one boundary component. Ascending states <strong>of</strong> ˜Gwhich contain one <strong>of</strong> α ±ε come in pairs S ∪ α ε <strong>and</strong> S ∪ α −ε with opposite signs,thus cancel out in A n,m ( ˜G). HenceA n,m (G) = A n,m ( ˜G).Let G <strong>and</strong> ˜G be two Gauss diagrams that differ by an application <strong>of</strong> Ω 3 , asshown in Figure 4.4 (G is on the left <strong>and</strong> ˜G is on the right).Then there is a bijective correspondence between ascending states <strong>of</strong> G <strong>and</strong>˜G. This correspondence preserves the signs <strong>and</strong> the combina<strong>to</strong>rics <strong>of</strong> the orderin which the tracing enters <strong>and</strong> leaves the neighborhood <strong>of</strong> these arrows. Thetable below summarizes this correspondence.69


112112323321323323112112112323321323323112211211113232313233223211211113232313233223For a better underst<strong>and</strong>ing <strong>of</strong> this table, let us explain one <strong>of</strong> the cases indetails. Denote the <strong>to</strong>p, left, <strong>and</strong> right arrows in the fragment by α t , α l , <strong>and</strong>α r respectively.Consider a state S ∪α l ∪α r <strong>of</strong> G which contains two arrows <strong>of</strong> the fragment.The order <strong>of</strong> tracing the fragment depends on S. Only two orders <strong>of</strong> tracingmay give an ascending state:1in1out1in1outStates <strong>of</strong> G :2out2in3in3ou<strong>to</strong>r3out3in2in2out1in1out1in1outStates <strong>of</strong> ˜G :2out2in3in3ou<strong>to</strong>r3out3in2in2outHere the three consecutive entries <strong>and</strong> exits from the fragment are indicatedby 1in, 1out, 2in, 2out, 3in, 3out. In the first case, the corresponding state <strong>of</strong>˜G is S ∪ α t ∪ α r . Note that the pattern <strong>of</strong> entries <strong>and</strong> exits from the fragment70


is indeed the same as in G. In the second case, the corresponding state <strong>of</strong> ˜G isS ∪ α t ∪ α l . The pattern <strong>of</strong> entries <strong>and</strong> exits is again the same as in G.4.2.2 Properties <strong>of</strong> A n,m (G) <strong>and</strong> D n,m (G)Theorem 4.2.3. Let G + , G − <strong>and</strong> G 0 be Gauss diagrams which differ only inthe fragment shown in Figure 4.9. Then∇ asc (G + ) − ∇ asc (G − ) = z∇ asc (G 0 ) ∇ des (G + ) − ∇ des (G − ) = z∇ des (G 0 ).(4.2)+ −Figure 4.9: A Conway triple <strong>of</strong> Gauss diagramsPro<strong>of</strong>. Again we will prove this theorem for ∇ asc . Denote by m <strong>and</strong> m 0 thenumber <strong>of</strong> circles in G ± <strong>and</strong> G 0 , respectively. It is enough <strong>to</strong> prove that forevery n ∈ NA n,m (G + ) − A n,m (G − ) = A n−1,m0 (G 0 ) (4.3)Denote the arrows <strong>of</strong> G + <strong>and</strong> G − appearing in the fragment in Figure 4.9 byα + <strong>and</strong> α − , respectively. The pro<strong>of</strong> <strong>of</strong> the skein relation is the same as in [15].All ascending states <strong>of</strong> G ± which do not contain α ± cancel out in (4.3) in pairs.Each ascending state S ∪ α ± <strong>of</strong> G ± corresponds <strong>to</strong> a unique ascending state S<strong>of</strong> G 0 , <strong>and</strong> vice versa: if S is an ascending state <strong>of</strong> G 0 , then (depending on theorder <strong>of</strong> the fragments in the tracing) exactly one <strong>of</strong> S ∪ α + <strong>and</strong> S ∪ α − willgive an ascending state on either G + or G − .It is easy <strong>to</strong> see that both A n,m (G) <strong>and</strong> D n,m (G) depend on the position<strong>of</strong> the base point when G is a Gauss diagram associated with a virtual linkdiagram. Let G <strong>and</strong> Ĝ be two Gauss diagrams shown in Figure 4.10. ThenA 2,1 (G) = 1, A 2,1 (Ĝ) = 0, D 2,1(G) = 0, D 2,1 (Ĝ) = 1.However, for classical links this is not the case. Our next theorem statesthat in the case <strong>of</strong> classical links, both ∇ asc (G) <strong>and</strong> ∇ des (G) are independen<strong>to</strong>f the position <strong>of</strong> the base point.71


G =+ +G^=+ +Figure 4.10: Dependence on a basepointTheorem 4.2.4. Let G be a based Gauss diagram <strong>of</strong> an m-component classicallink L. Both ∇ asc (G) <strong>and</strong> ∇ des (G) are independent <strong>of</strong> the position <strong>of</strong> the basepoint.Pro<strong>of</strong>. We will prove the independence <strong>of</strong> A n,m (G) <strong>of</strong> the position <strong>of</strong> the basepoint, the pro<strong>of</strong> for D n,m (G) is the same. We prove this statement by inductionon the number <strong>of</strong> arrows in G.If G has no arrows, there is nothing <strong>to</strong> prove. Now let us assume that thestatement holds for any (classical) Gauss diagram with less than k arrows, <strong>and</strong>let G be a Gauss diagram with k arrows. If k < n, then A n,m (G) = 0 <strong>and</strong> weare done, so we may assume that k ≥ n. Suppose that the base point lies onthe i-th component <strong>of</strong> G. We should prove that we may move the base pointacross any arrowhead or arrowtail on the i-th component, <strong>and</strong> <strong>to</strong> shift it <strong>to</strong> anyother, say, j-th, component. Denote by G 1 , . . . , G 7 Gauss diagrams which differonly in a fragment which looks likei j i j i j i j i j i j i jG 1 G 2 G 3 G 4 G 5 G 6 G 7respectively. It suffices <strong>to</strong> prove that we have:A n,m (G 1 ) = A n,m (G 2 ) , A n,m (G 3 ) = A n,m (G 4 ) , A n,m (G 3 ) = A n,m (G 5 ) (4.4)Denote by α the arrow appearing in the fragment above. The first equalityis immediate; indeed, in G 1 <strong>and</strong> G 2 there are no ascending states with oneboundary component which contain α (<strong>and</strong> all other ascending states are in abijective correspondence).72


To prove the second equation in (4.4), note that there is a bijection betweenascending states <strong>of</strong> A n,m (G 3 ) <strong>and</strong> A n,m (G 4 ) which do not contain α; the remainingascending states <strong>of</strong> G 3 <strong>and</strong> G 4 look exactly like the ones <strong>of</strong> G 6 <strong>and</strong> G 7 ,respectively; thus we haveA n,m (G 3 ) − A n,m (G 4 ) = A n−1,m0 (G 6 ) − A n−1,m0 (G 7 ) = 0,where m 0 = m±1 is the number <strong>of</strong> circles in G 6 <strong>and</strong> G 7 <strong>and</strong> the second equalityholds by the induction hypothesis.The pro<strong>of</strong> <strong>of</strong> the last equality in (4.4) is more complicated. We will use aninner induction on the number r <strong>of</strong> arrows which have only <strong>their</strong> arrowtail onthe i-th component.If r = 0, then A n,m (G 3 ) = A n,m (G 5 ) = 0. Indeed, for r = 0 there are noascending states in G 5 (since we cannot reach the i-th circle), so A n,m (G 5 ) = 0.Also, since r = 0, i-th component <strong>of</strong> the link is under all other components, sowe may move it apart by a finite sequence <strong>of</strong> Reidemeister moves Ω 2 <strong>and</strong> Ω 3applied away from the base point, converting G 3 in<strong>to</strong> a Gauss diagram withan isolated i-th component (here we use the fact that G 3 is associated with aclassical link diagram); thus A n,m (G 3 ) = 0 by Theorem 4.2.2.Let’s establish the step <strong>of</strong> induction. On both G 3 <strong>and</strong> G 5 pick the samearrow, which has only its arrowtail on the i-th component, <strong>and</strong> apply the skeinrelation <strong>of</strong> Theorem 4.2.3 <strong>to</strong> simplify the corresponding Gauss diagrams. Diagramson the right-h<strong>and</strong> side <strong>of</strong> the skein relation have less than k crossings,so the right-h<strong>and</strong> sides are equal by the induction on k; the remaining terms inthe left-h<strong>and</strong> side are also equal by induction on r.Corollary 4.2.5. Let G be a Gauss diagram <strong>of</strong> an m-component classical linkL. Then∇ asc (G) = ∇ des (G) = ∇(L),i.e. for every n ≥ 0A n,m (G) = D n,m (G) = c n (L). (4.5)Pro<strong>of</strong>. By Theorems 4.2.2 – 4.2.4, A n,m (G) <strong>and</strong> D n,m (G) are link <strong>invariants</strong>which satisfy the same skein relation as c n (L). It remains <strong>to</strong> compare <strong>their</strong>normalization, i.e. <strong>their</strong> values on the unknot O. A st<strong>and</strong>ard Gauss diagramG(O) <strong>of</strong> an unknot consists <strong>of</strong> one circle with no arrows. ThusD 0,1 (G(O)) = A 0,1 (G(O)) = 1, <strong>and</strong> D n,m (G(O)) = A n,m (G(O)) = 0 otherwise.Hence A n,m (G) <strong>and</strong> D n,m (G) have the same normalization as c n (L) <strong>and</strong> thecorollary follows.73


Example 4.2.6. Consider a 3-component link L <strong>and</strong> the Gauss diagram G <strong>of</strong>L shown below:4−−−−1 23The only ascending state <strong>of</strong> G is {3, 4}. It sign is +. Thus c 2 (L) = 1 <strong>and</strong>c n (L) = 0 for all n ≠ 2, so ∇(L) = z 2 .4.2.3 Alex<strong>and</strong>er-Conway polynomials <strong>of</strong> long virtual linksIn this subsection we study properties <strong>of</strong> the polynomials ∇ asc <strong>and</strong> ∇ des for longvirtual links, <strong>and</strong> compare them with other existing <strong>constructions</strong>.Let L be a classical or long virtual link <strong>and</strong> G be any Gauss diagram <strong>of</strong>L. Polynomials ∇ asc (L) := ∇ asc (G) <strong>and</strong> ∇ des (L) := ∇ des (G) were defined in[15], but the pro<strong>of</strong> that they are well defined for classical links with more than 2components <strong>and</strong> for long virtual links was not presented. By Theorem 4.2.2 <strong>and</strong>Corollary 4.2.5 ∇ asc (L) <strong>and</strong> ∇ des (L) are <strong>invariants</strong> <strong>of</strong> L, <strong>and</strong> if L is a classicallink∇ asc (L) = ∇ des (L) = ∇(L).Note that for long virtual links it may happen that∇ asc (L) ≠ ∇ des (L).For example, let G be a Gauss diagram <strong>of</strong> the long virtual Hopf link L shownin Figure 4.2. Then ∇ asc (L) = z, but ∇ des (L) = 0. We denote by∇ asc−des (L) := ∇ asc (L) − ∇ des (L).This polynomial vanishes on classical links but, as we will see below, may beused <strong>to</strong> distinguish virtual links from classical links.Let D be a diagram <strong>of</strong> a (long) virtual link L <strong>and</strong> G its corresponding Gaussdiagram. Pick a classical crossing on D. A move on D <strong>and</strong> the correspondingmove on G shown in Figure 4.11 is called the virtualization move.Theorem 4.2.7. Let L <strong>and</strong> L 1 be long virtual links. Then the following holds.a) ∇ asc (L#L 1 ) = ∇ asc (L)∇ asc (L 1 ) , ∇ des (L#L 1 ) = ∇ des (L)∇ des (L 1 ).74


εεFigure 4.11: The virtualization moveb) Non-trivial coefficients <strong>of</strong> ∇ des <strong>and</strong> ∇ asc are not invariant under the virtualizationmovec) Coefficients <strong>of</strong> ∇ des (L) <strong>and</strong> ∇ asc (L) are Vassiliev <strong>invariants</strong> in the sense<strong>of</strong> both GPV [30] <strong>and</strong> Kauffman [35].Pro<strong>of</strong>. The pro<strong>of</strong> <strong>of</strong> Part a) is straightforward <strong>and</strong> follows from the definition<strong>of</strong> ∇ asc <strong>and</strong> ∇ des .Part b). Consider even n first. Let n = 2k <strong>and</strong> let G <strong>and</strong> G 1 be Gaussdiagrams <strong>of</strong> long virtual knots shown in Figure 4.12a <strong>and</strong> 4.12b respectively.Note that G <strong>and</strong> G 1 differ by an application <strong>of</strong> the virtualization move. BothG <strong>and</strong> G 1 have n + 1 arrows. It is easy <strong>to</strong> check thatA n,1 (G) = −1 D n,1 (G) = 1, but A n,1 (G 1 ) = D n,1 (G 1 ) = 0.To prove the statement for odd n, add a Hopf-linked unknot <strong>to</strong> the abovediagrams.Part c) It is enough <strong>to</strong> prove that A n,m (G) <strong>and</strong> D n,m (G) are GPV finite type<strong>invariants</strong>, because any GPV finite type invariant is au<strong>to</strong>matically <strong>of</strong> Kauffmanfinite type. But [30, page 12] implies, that any invariant given by an arrowdiagram formula with n arrows is GPV finite type <strong>of</strong> degree n.Remark 4.2.8. It follows from [19, Theorem 1], that Part b) in Theorem 4.2.7follows from Part c).Let K T be a virtual knot with a virtual diagram shown in Figure 4.13. Thisknot is called the Kishino knot. It has attracted attention for its remarkableproperty that it is a connected sum <strong>of</strong> two diagrams <strong>of</strong> the trivial knot; it hastrivial Jones polynomial, Z ′ K T(t, y) = 0 (for the definition <strong>of</strong> Z ′ see [58]), <strong>and</strong>75


−−2k2k−−++abFigure 4.12: Diagrams <strong>of</strong> long virtual knots that differ by an application <strong>of</strong> thevirtualization movethe virtual knot group <strong>of</strong> K T is isomorphic <strong>to</strong> Z, see [38]. It was first proved <strong>to</strong>be non-classical in [38]. We show that K T is non-classical using the polynomials∇ asc , ∇ des <strong>and</strong> ∇ asc−des .Figure 4.13: A diagram <strong>of</strong> the Kishino knotProposition 4.2.9. Polynomials ∇ asc , ∇ des <strong>and</strong> ∇ asc−des detect the fact thatK T is a non-classical knot.Pro<strong>of</strong>. Recall that for any Gauss diagram G <strong>of</strong> a classical knot all these polynomialsare independent <strong>of</strong> the position <strong>of</strong> the base point. Consider two Gaussdiagrams G <strong>and</strong> Ĝ <strong>of</strong> K T which differ only by the position <strong>of</strong> the base point,see Figure 4.14.We have∇ asc (G) = 1 − 2z 2 + z 4 ∇ des (G) = 1 ∇ asc−des (G) = −2z 2 + z 4 but∇ asc (Ĝ) = 1 ∇ des(Ĝ) = 1 − 2z2 + z 4 ∇ asc−des (Ĝ) = 2z2 − z 4 .76


^G = G−−=−+ ++ +−Figure 4.14: Gauss diagrams <strong>of</strong> the Kishino knotComparison with other <strong>constructions</strong> <strong>of</strong> Alex<strong>and</strong>er-Conwaypolynomials <strong>of</strong> virtual linksIn [58] Sawollek associated <strong>to</strong> every link diagram D <strong>of</strong> a virtual link L a Laurentpolynomial Z D ′ (t, y) in two variables t, y. He proved that Z′ D (t, y) is an invarian<strong>to</strong>f virtual links up <strong>to</strong> multiplication by powers <strong>of</strong> t ±1 , <strong>and</strong> that it vanishes onclassical links. He also showed that Z D ′ (t, y) satisfies the following skein relation:Z ′ D +(t, y) − Z ′ D −(t, y) = (t −1 − t)Z ′ D 0(t, y).It is obvious that both ∇ asc (L) <strong>and</strong> ∇ des (L) are crucially different from Z L ′ (t, y)because both <strong>of</strong> them do not vanish on classical links, but one can suspectthat ∇ asc−des (L) coincides with Z L ′ (t, y) after a possible renormalization <strong>and</strong> achange <strong>of</strong> variables z = t −1 − t. Sawollek proved the following theorem:Theorem 4.2.10 ([58]). Let D, D 1 , D 2 be virtual link diagrams <strong>and</strong> let D 1 ⊔D 2denote the disconnected sum <strong>of</strong> the diagrams D 1 <strong>and</strong> D 2 . ThenZ ′ D 1 ⊔D 2(t, y) = Z ′ D 1(t, y)Z ′ D 2(t, y).Note that ∇ asc−des (L ⊔ L) = 0 for any long virtual link L, but Z L⊔L ′ (t, y) =(Z L ′ (t, y))2 . Thus ∇ asc−des (L) is also drastically different from Z L ′ (t, y).Another generalizations <strong>of</strong> Alex<strong>and</strong>er polynomials <strong>to</strong> virtual links are derivedfrom the virtual <strong>and</strong> extended virtual link groups, see [58] <strong>and</strong> [60, 61]respectively.1. Following [58] we denote by ∆ L (t) a polynomial which is derived from thevirtual link group <strong>of</strong> a link L. It is well defined up <strong>to</strong> sign <strong>and</strong> multiplicationby powers <strong>of</strong> t ±1 . For every virtual link diagram D the associated polynomialis denoted by ∆ D (t). In contrast <strong>to</strong> the classical Alex<strong>and</strong>er polynomial, theAlex<strong>and</strong>er polynomial <strong>of</strong> [58] for virtual links does not satisfy any linear skeinrelation as stated in the next theorem:77


Theorem 4.2.11 ([58]). For any normalization A D (t) <strong>of</strong> the polynomial ∆ D (t),i.e., A D (t) = ε D t n D∆ D (t) with some ε D ∈ {−1, 1} <strong>and</strong> n D ∈ Z, the equationp 1 (t)A D+ (t) + p 2 (t)A D− (t) + p 3 (t)A D0 (t) = 0 with p 1 (t), p 2 (t), p 3 (t) ∈ Z[t ±1 ] hasonly the trivial solution p 1 (t) = p 2 (t) = p 3 (t) = 0.Since ∇ asc , ∇ des <strong>and</strong> ∇ asc−des satisfy the Conway skein relation, it followsthat all <strong>of</strong> them are crucially different from ∆ D (t) <strong>of</strong> [58].2. Let L be a virtual link. The polynomial ∆ 1 (L) <strong>of</strong> [60, 61] is a polynomialin variables v, u <strong>and</strong> is well defined up <strong>to</strong> multiplication by powers <strong>of</strong> (uv) ±1 .If L is a classical link, then ∆ 1 (L) is equal <strong>to</strong> the Alex<strong>and</strong>er polynomial in thevariable uv. It follows that ∇ asc−des is different from ∆ 1 , because ∆ 1 is notidentically zero on the family <strong>of</strong> classical links.Given a virtual link L, we denote by L ∗ the mirror image <strong>of</strong> L, i.e. a linkobtained by inverting the sign <strong>of</strong> each classical crossing in a diagram <strong>of</strong> L. Thefollowing corollary was proved in [61].Corollary 4.2.12 ([61], Corollary 5.2). Let L be a virtual m-component link.Then∆ 1 (L)(u, v) = (−1) m ∆ 1 (L ∗ )(v, u)up <strong>to</strong> multiplication by powers <strong>of</strong> (uv) ±1 .In particular, for any virtual knot ∆ 1 (K)(u, v) = −∆ 1 (K ∗ )(v, u).Consider a mirror pair <strong>of</strong> long virtual knots K <strong>and</strong> K ∗ with Gauss diagramsG <strong>and</strong> G ∗ shown in Figure 4.15.G= G * =+ + − −Figure 4.15: A mirror pair <strong>of</strong> Gauss diagramsThen ∇ asc (K) = ∇ des (K ∗ ) = 1 + z 2 , ∇ asc (K ∗ ) = ∇ des (K) = 1. Thus both ∇ asc<strong>and</strong> ∇ des are crucially different from ∆ 1 (L) in [60, 61].Another way <strong>to</strong> see that both ∇ asc <strong>and</strong> ∇ des are different from ∆ 1 is byfinding a virtual knot K, such that both ∇ asc <strong>and</strong> ∇ des detect that this knotis non-classical, but ∆ 1 (K) = 1 (so ∆ 1 does not distinguish this knot from theunknot). An example <strong>of</strong> such a knot K, <strong>to</strong>gether with a pair G <strong>and</strong> Ĝ <strong>of</strong> its78


G=−−−−G^=−−−−Figure 4.16: A virtual knot <strong>and</strong> two <strong>of</strong> its based Gauss diagramsGauss diagrams which differ by the position <strong>of</strong> the base point, is given in Figure4.16. It was shown in [61] that ∆ 1 (K) = 1.We have∇ asc (G) = 1 + z 2 ∇ des (G) = 1 + z 2 but∇ asc (Ĝ) = 1 + 2z2 + z 4 ∇ des (Ĝ) = 1.It follows that both ∇ asc <strong>and</strong> ∇ des show that K is non-classical.Finally, another generalization <strong>of</strong> the Alex<strong>and</strong>er polynomial (related <strong>to</strong> thepolynomial Z ′ <strong>of</strong> [58]) <strong>to</strong> long virtual knots was presented in [1]. It is a Laurentpolynomial ζ in a variable t over the following ring T = Z[p, p −1 , q, q −1 ]/((p −1)(p − q), (q − 1)(p − q)). This polynomial also vanishes on classical knots <strong>and</strong>thus ζ significantly differs from ∇ asc <strong>and</strong> ∇ des .Question. Is it possible <strong>to</strong> derive ∇ asc−des from ζ?4.3 Counting surfaces with two boundarycomponentsIn this section we present a new infinite family <strong>of</strong> Gauss diagram formulas,which correspond <strong>to</strong> counting <strong>of</strong> orientable surfaces with two boundary components.At the end <strong>of</strong> this section we identify the resulting <strong>invariants</strong> withcertain derivatives <strong>of</strong> the HOMFLYPT polynomial.4.3.1 Link <strong>invariants</strong> <strong>and</strong> diagrams with two boundarycomponentsIn this subsection we define <strong>invariants</strong> <strong>of</strong> classical links using ascending <strong>and</strong>descending arrow diagrams with two boundary components.79


Recall that for every Gauss diagram G we defined notions <strong>of</strong> ascending <strong>and</strong>descending separating states <strong>of</strong> G, see Subsection 4.1.5. Also, for every ascending(respectively descending) separating state S <strong>of</strong> G <strong>and</strong> for an arrow diagramA ∈ A 2 n,m (respectively A ∈ Dn,m) 2 we defined, in the same subsection, a notion<strong>of</strong> S-admissible pairing ⟨A, G⟩ S . Note that every ascending (respectivelydescending) separating state S <strong>of</strong> G defines two Gauss diagrams G ′ S <strong>and</strong> G′′ S asfollows: G ′ S (respectively G′′ S ) consists <strong>of</strong> all circles <strong>of</strong> G S labeled by 1 (respectivelyby 2), <strong>and</strong> its arrows are arrows <strong>of</strong> G with both ends on these circles. Allarrows with ends on circles <strong>of</strong> G S with different labels are removed. The basepoint on G ′ is the base point ∗ <strong>of</strong> G. The base point on G ′′ is placed near thefirst arrow in S which we encounter as we walk on G starting from ∗.If G is a Gauss diagram <strong>of</strong> a classical link L, then G ′ S <strong>and</strong> G′′ S correspond <strong>to</strong>classical links L ′ S <strong>and</strong> L′′ S , which are defined as follows. We smooth all crossingswhich correspond <strong>to</strong> arrows in S, as shown below:We obtain a diagram <strong>of</strong> a smoothed link L S with labeling <strong>of</strong> components inducedfrom the labeling <strong>of</strong> circles <strong>of</strong> G S . Denote by L ′ S <strong>and</strong> L′′ S sublinks which consis<strong>to</strong>f components labeled by 1 <strong>and</strong> 2 respectively.It follows from Corollary 4.2.5 that for every n ≥ 0 we haveA n,m ′(G ′ S) = D n,m ′(G ′ S) = c n (L ′ S) <strong>and</strong> A n,m ′′(G ′′ S) = D n,m ′′(G ′′ S) = c n (L ′′ S),where m ′ <strong>and</strong> m ′′ are the number <strong>of</strong> components <strong>of</strong> L ′ S <strong>and</strong> L′′ S respectively.Using this <strong>to</strong>gether with the definition <strong>of</strong> A 2 n,m(G) S <strong>and</strong> D 2 n,m(G) S in Subsection4.1.5 we getLemma 4.3.1. Let G be a Gauss diagram <strong>of</strong> an m-component link L. Thenfor every n ≥ 0 <strong>and</strong> an ascending (respectively descending) separating state S<strong>of</strong> G we haven−|S|∑A 2 n,m(G) S = sign(S) c i (L ′ S)c n−|S|−i (L ′′ S)i=0n−|S|∑Dn,m(G) 2 S = sign(S) c i (L ′ S)c n−|S|−i (L ′′ S)i=0<strong>and</strong>respectively.Here |S| is the number <strong>of</strong> arrows in S <strong>and</strong> sign(S) = ∏ α∈S sign(α).80


Summing over all ascending (descending) separating states S, we obtainCorollary 4.3.2. Let G be any Gauss diagram <strong>of</strong> an m-component link L.Then for every n ≥ 0 we haveA 2 n,m(G) =D 2 n,m(G) =n∑∑k=1 S,|S|=kn∑∑k=1 S,|S|=k∑n−ksign(S) c i (L ′ S)c n−k−i (L ′′ S)i=0∑n−ksign(S) c i (L ′ S)c n−k−i (L ′′ S)where the second summation is over all ascending <strong>and</strong> descending separatingstates S <strong>of</strong> G respectively.It turns out that both A 2 n,m(G) <strong>and</strong> D 2 n,m(G) are invariant under Ω 2 <strong>and</strong> Ω 3 :Theorem 4.3.3. Let G be any Gauss diagram <strong>of</strong> an m-component link L. ThenA 2 n,m(G) <strong>and</strong> D 2 n,m(G) are invariant under Reidemeister moves Ω 2 <strong>and</strong> Ω 3 whichdo not involve the base point.Pro<strong>of</strong>. We will prove the invariance <strong>of</strong> A 2 n,m(G); the pro<strong>of</strong> for D 2 n,m(G) is thesame.Let G <strong>and</strong> ˜G be two Gauss diagrams that differ by an application <strong>of</strong> Ω 2 ,so that ˜G has two additional arrows α ε <strong>and</strong> α −ε , see Figure 4.4. Ascendingstates <strong>of</strong> G are in bijective correspondence with ascending states <strong>of</strong> ˜G whichdo not contain α ±ε . Note that α ε <strong>and</strong> α −ε can not be both in the image <strong>of</strong>ϕ : A → G with A ∈ A 2 n,m, because A is an ascending diagram with twoboundary components. Ascending states <strong>of</strong> ˜G which contain one <strong>of</strong> α ±ε comein pairs S ∪ α ε <strong>and</strong> S ∪ α −ε with opposite signs, thus cancel out in A 2 n,m( ˜G).HenceA 2 n,m(G) = A 2 n,m( ˜G).Now, let G <strong>and</strong> ˜G be two Gauss diagrams that differ by an application <strong>of</strong>Ω 3 , see Figure 4.4 (G is on the left <strong>and</strong> ˜G is on the right). Denote the <strong>to</strong>p,left, <strong>and</strong> right arrows in the fragment by α t , α l , <strong>and</strong> α r respectively. There is abijective correspondence between ascending separating states <strong>of</strong> G <strong>and</strong> ˜G, suchthat none <strong>of</strong> the arrows α r , α l <strong>and</strong> α t belong <strong>to</strong> these states. Indeed, we mayidentify separating states <strong>of</strong> G <strong>and</strong> ˜G which have the same arrows <strong>and</strong> the samelabeling <strong>of</strong> arcs. For any such separating state S we have A 2 n,m(G) S = A 2 n,m( ˜G) Sby Lemma 4.3.1 <strong>and</strong> Theorem 4.2.2.81i=0


122GG ~2 1 12 21111SL S~ S12 1 111~SLL SL ~ SFigure 4.17: Identifying ascending separating states containing α lAn ascending separating state <strong>of</strong> G or ˜G may contain either exactly onearrow <strong>of</strong> the fragment i.e. α r or α l or α t , or it may contain both arrows α r <strong>and</strong>α l . There is a bijective correspondence between ascending separating states <strong>of</strong>G <strong>and</strong> ˜G which contain α l . Two possible cases <strong>of</strong> this correspondence (whichdiffer by the labeling) are shown in Figure 4.17 <strong>and</strong> Figure 4.18.G1222 11222SL SL SG ~12 1 2222 21~SL ~ SL ~ SFigure 4.18: Identifying other ascending separating states containing α lIn both cases, links L ′ S <strong>and</strong> L′′ S constructed from G <strong>and</strong> ˜G are iso<strong>to</strong>pic,thus by Lemma 4.3.1 A 2 n,m(G) S = A 2 n,m( ˜G) ˜S.The situation with ascendingseparating states which contain α r is completely similar <strong>and</strong> is omitted.The correspondence <strong>of</strong> ascending separating states which contain α l ∪ α r or82


GG1 11 12 22 2 21 12 22 22 1 2S t L t L tS lr L lr L lrG ~1 1222 2 2 2 2~S t~L t~L tFigure 4.19: Comparison <strong>of</strong> ascending separating states <strong>of</strong> G <strong>and</strong> ˜Gα t is more complicated. One <strong>of</strong> the two possible cases is summarized in Figure4.19. Links ˜L ′ t, L ′ t <strong>and</strong> L ′ lr are iso<strong>to</strong>pic, thusc i ( ˜L ′ t) = c i (L ′ t) = c i (L ′ lr).Moreover, for c i (˜L ′′t ) we have:( ) ((c i + + = c i + −)+c i−1+)= c i( )+c i−1(+)The first equality is the skein relation <strong>of</strong> Theorem 4.2.3, <strong>and</strong> the second equalityholds by the invariance <strong>of</strong> c i under Ω 2 . Hencec i (˜L ′′t ) = c i (L ′′t ) + c i−1 (L ′′lr).Denote by k the number <strong>of</strong> arrows in S t . Note that the number <strong>of</strong> arrows in ˜S t<strong>and</strong> S lr is k <strong>and</strong> k + 1, respectively. Thus∑n−ki=0∑n−kc i (˜L ′ t)c n−k−i (˜L ′′t ) =i=0n−k−1∑c i (L ′ t)c n−k−i (L ′′t ) +83i=0c i (L ′ lr)c n−k−i−1 (L ′′lr).


G1 11 12 2S t1 1 1L tL tG ~ G ~1 1222 21 1 112 1 1 121 2 1~S t~S lr~L t~L lr~L t~L lrFigure 4.20: Comparison <strong>of</strong> other ascending separating states <strong>of</strong> G <strong>and</strong> ˜GNote that sign( ˜S t ) = sign(S t ) = sign(S lr ), thus by Lemma 4.3.1A 2 n,m( ˜G) ˜St= A 2 n,m(G) St + A 2 n,m(G) Slr .The second possible case (which differs by labeling) is shown in Figure 4.20.Abusing the notation we again denote the corresponding ascending separatingstates by S t , ˜S t , ˜S lr . Links L ′′ ′′ ′′t , ˜L t <strong>and</strong> ˜L lr are iso<strong>to</strong>pic. Applying the skeinrelation for c i (L ′ t) similarly <strong>to</strong> the above, in this case we getA 2 n,m(G) St = A 2 n,m( ˜G) ˜St+ A 2 n,m( ˜G) ˜Slr.Our next step is <strong>to</strong> study the behavior <strong>of</strong> A 2 n,m(G) <strong>and</strong> D 2 n,m(G) under anapplication <strong>of</strong> Reidemeister move Ω 1 . Both A 2 n,m(G) <strong>and</strong> D 2 n,m(G) change underΩ 1 , see Example 4.3.4.Example 4.3.4. Let G, G 1 <strong>and</strong> G 2 be Gauss diagrams <strong>of</strong> an unknot shown inFigures 4.21a, 4.21b <strong>and</strong> 4.21c respectively. Then D 2 1,1(G) = 0, butD 2 1,1(G 1 ) = 1; <strong>and</strong> A 2 1,1(G) = 0, but A 2 1,1(G 2 ) = −1.84


+ −a b cFigure 4.21: Gauss diagrams G, G 1 <strong>and</strong> G 2However, this problem is easy <strong>to</strong> solve. Denote by<strong>and</strong> letAD n,m (G) := A 2 n,m(G) + D 2 n,m(G)I n,m (G) := AD n,m (G) − w(G)c n−1 (L),where w(G) is the writhe <strong>of</strong> G, i.e. the sum <strong>of</strong> signs <strong>of</strong> all arrows in G.Theorem 4.3.5. Let G be a Gauss diagram <strong>of</strong> a long m-component link L.Then I n,m (G) is an invariant <strong>of</strong> an underlying link L, i.e. is independent <strong>of</strong> achoice <strong>of</strong> G.Pro<strong>of</strong>. By Theorem 4.3.3, it remains <strong>to</strong> prove the invariance <strong>of</strong> I n,m under Ω 1(applied away from the base point). Let ˜G <strong>and</strong> G be two Gauss diagrams whichare related by an application <strong>of</strong> Ω 1 , such that ˜G contains a new isolated arrow α.Then α contributes either a new ascending or a new descending separating state{a}, depending on whether we meet its head or tail first on the passage from thebase point. Contribution <strong>of</strong> this state <strong>to</strong> AD n,m ( ˜G) is either sign(α)A n−1,m (G),or sign(α)D n−1,m (G); butsign(α)A n−1,m (G) = sign(α)c n−1 (L) = sign(α)D n−1,m (G)by Corollary 4.2.5, thusI n,m ( ˜G) − I n,m (G) = (sign(α) − w( ˜G) + w(G))c n−1 (L).It remains <strong>to</strong> note that w( ˜G) − w(G) = sign(α).Our next step is <strong>to</strong> study dependence <strong>of</strong> A 2 n,m(G) <strong>and</strong> D 2 n,m(G) on the position<strong>of</strong> the base point. The example below shows that each <strong>of</strong> them dependson the base point.85


Example 4.3.6. Let G <strong>and</strong> Ĝ be two Gauss diagrams <strong>of</strong> the right-h<strong>and</strong>edtrefoil shown below. Then A 2 3,1(G) = 0 <strong>and</strong> D3,1(G) 2 2= 1, but A3,1(Ĝ) = 1 <strong>and</strong>D 2 3,1(Ĝ) = 0. + +G=+G^ =++ +It turns out, however, that the sum A 2 n,m(G) + Dn,m(G) 2 = AD n,m (G) doesnot depend on the base point. Indeed, let G <strong>and</strong> Ĝ be two Gauss diagramswhich differ only by a position <strong>of</strong> <strong>their</strong> base points. Let Ŝ be an ascendingseparating state <strong>of</strong> Ĝ. If an arc which contains the base point ∗ <strong>of</strong> G is labeledby 1, then S = Ŝ is an ascending separating state <strong>of</strong> G <strong>and</strong> by Lemma 4.3.1 wehave A 2 n,m(G) S = A 2 n,m(Ĝ) Ŝ. If an arc which contains the base point ∗ <strong>of</strong> G islabeled by 2, we consider a descending separating state S <strong>of</strong> G which has thesame arrows as Ŝ, but opposite labels. By Lemma 4.3.1, D2 n,m(G) S = A 2 n,m(Ĝ) Ŝ .We repeat this process, replacing ascending separating states with descending<strong>and</strong> A with D. Summing up by separating states, in view <strong>of</strong> Corollary 4.3.2 weobtain the followingTheorem 4.3.7. Let G be a based Gauss diagram <strong>of</strong> an m-component link L.Then AD n,m (G) is independent <strong>of</strong> the position <strong>of</strong> the base point.It has two important corollaries.Corollary 4.3.8. Let G be any Gauss diagram <strong>of</strong> an m-component framed linkL. Then AD n (L) = AD n,m (G) is an invariant <strong>of</strong> an underlying framed link, i.edoes not depend on G.Pro<strong>of</strong>. It is easy <strong>to</strong> see that, if Gauss diagrams G <strong>and</strong> ˜G differ by Ω F 1 move,then AD n,m (G) = AD n,m ( ˜G). Now the pro<strong>of</strong> follows directly from Theorems4.3.3 <strong>and</strong> 4.3.7.Corollary 4.3.9. Let G be any Gauss diagram <strong>of</strong> an m-component link L. ThenI n (L) := I n,m (G) is an invariant <strong>of</strong> an underlying link L, i.e. is independent <strong>of</strong>a choice <strong>of</strong> G.4.3.2 Properties <strong>of</strong> I nIn this subsection we establish the skein relation for I n,m . Then we identify I n,mwith coefficients <strong>of</strong> a certain polynomial, which is derived from the HOMFLYPTpolynomial.86


Skein relationIn this part we establish the skein relation for AD n,m (G).Theorem 4.3.10. Let L + , L − , L 0 be a Conway triple <strong>of</strong> links with the correspondingConway triple G + , G − , G 0 <strong>of</strong> Gauss diagrams, see Figures 1.4 <strong>and</strong>4.9. Denote the number <strong>of</strong> circles <strong>of</strong> G ± <strong>and</strong> G 0 by m <strong>and</strong> m 0 , respectively.ThenAD n,m (G + ) − AD n,m (G − ) =⎧⎪⎨ AD n−1,m0 (G 0 ) , if m 0 = m − 1⎪⎩ AD n−1,m0 (G 0 ) + ∑ ∑c i (L ′ )c n−i−1 (L 0 L ′ ) if m 0 = m + 1 (4.6)n−1L ′ ⊂L 0 i=0where the summation is over all sublinks L ′ <strong>of</strong> L 0 which contain exactly onenew component resulting from the smoothing.Pro<strong>of</strong>. Denote the arrows <strong>of</strong> G + <strong>and</strong> G − appearing in Figure 4.9 by α + <strong>and</strong> α − ,respectively.Let us look at labels <strong>of</strong> ascending separating states <strong>of</strong> G ± <strong>and</strong> G 0 on four arcs<strong>of</strong> the shown fragment. If labels <strong>of</strong> all four arcs are the same, we may identifystates <strong>of</strong> G ± <strong>and</strong> G 0 with the same arrows <strong>and</strong> labels <strong>of</strong> arcs, see Figure 4.22a.Lemma 4.3.1 <strong>and</strong> Theorem 4.2.3 imply, that for every such state SA 2 n,m(G + ) S − A 2 n,m(G − ) S = A 2 n−1,m 0(G 0 ) S .If labels on two arcs near the head <strong>of</strong> α ± coincide, but differ from labels nearthe tail <strong>of</strong> α ± , by Lemma 4.3.1 we have A 2 n,m(G + ) S − A 2 n,m(G − ) S = 0 for anysuch state S <strong>of</strong> G ± , <strong>and</strong> there is no corresponding state <strong>of</strong> G 0 . See Figure 4.22b(for i ≠ j).There are two further cases when labels <strong>of</strong> two arcs near the head <strong>of</strong> α ± aredifferent. Such a state S <strong>of</strong> G 0 corresponds either <strong>to</strong> an ascending separatingstate S ∪α + <strong>of</strong> G + , or <strong>to</strong> an ascending separating state S ∪α − <strong>of</strong> G − , see Figure4.22c. By Lemma 4.3.1 we have A 2 n,m(G + ) S∪α+ = A 2 n−1,m 0(G 0 ) S in the first case<strong>and</strong> A 2 n,m(G − ) S∪α− = −A 2 n−1,m 0(G 0 ) S in the second case.If m 0 = m − 1, there are no other ascending separating states <strong>of</strong> any <strong>of</strong>the diagrams <strong>and</strong>, repeating this computation for descending separating states,summing over states <strong>and</strong> using Corollary 4.3.2, we obtain the first equality in(4.6).87


iiii+iiii+jj2 2 2 + 21 1 1 1iiaii_iiii_bjj1 1 1 _ 12 2 2 2cFigure 4.22: Correspondence <strong>of</strong> separating states <strong>of</strong> G 0 <strong>and</strong> G ±If m 0 = m + 1, both ends <strong>of</strong> α ± are on the same circle <strong>of</strong> G ± <strong>and</strong> there isan additional contribution <strong>to</strong> AD n,m (G ± ) <strong>of</strong> separating states S = {α ± } <strong>of</strong> G ± ,which contain only the arrow α ± (<strong>and</strong> some labeling <strong>of</strong> arcs) 1 . Such separatingstates correspond <strong>to</strong> labeling all circles <strong>of</strong> G 0 by 1, 2 so that the based circle islabeled by 1, <strong>and</strong> two new components <strong>of</strong> G 0 resulting from the smoothing havedifferent labels. Denote by L ′ the sublink labeled by 1. The case <strong>of</strong> descendingseparating states {α ± } is similar. By Corollary 4.3.2, the contribution <strong>of</strong> thesestates <strong>to</strong> AD n,m (G + ) − AD n,m (G − ) equals<strong>and</strong> the theorem follows.∑n−1L ′ ⊂L 0 i=0∑c i (L ′ )c n−i−1 (L 0 L ′ )Identification <strong>of</strong> the invariant I nIn this subsubsection we identify I n with certain derivatives <strong>of</strong> the HOMFLYPTpolynomial.Let P (L) be the HOMFLYPT polynomial <strong>of</strong> a link L. We denote by P ′ a(L)the first derivative <strong>of</strong> P (L) w.r.t. a. Then P ′ a(L)| a=1 is a polynomial in thevariable z. We denote by p n (L) the coefficient <strong>of</strong> z n in zP ′ a(L)| a=1 .Theorem 4.3.11. Let L be an m-component link. Then for every n ≥ 0I n (L) = p n (L) − C n (L), (4.7)1 These states have no counterpart in G 0 , since such a separating state <strong>of</strong> G 0 should beempty <strong>and</strong> the corresponding surface disconnected.88


where C n (L) is defined byC n (L) := ∑ L ′ ⊂Ln∑c i (L ′ )c n−i (L L ′ ),<strong>and</strong> the summation is over all proper sublinks L ′ <strong>of</strong> L.i=0Pro<strong>of</strong>. It is enough <strong>to</strong> show that I n (L)+C n (L) <strong>and</strong> p n (L) satisfy the same skeinrelation <strong>and</strong> take the same value on unlinks with any number <strong>of</strong> components.The skein relation for zP ′ a(L)| a=1 follows directly from the skein relationfor P (L). Differentiating this skein relation w.r.t. a, substituting a = 1, <strong>and</strong>multiplying by z we obtainzP ′ a(L + )| a=1 − zP ′ a(L − )| a=1 + zP (L + )| a=1 + zP (L − )| a=1 = z 2 P ′ a(L 0 )| a=1 .Note that P (L)| a=1 = ∇(L) is the Conway polynomial <strong>of</strong> L. ThuszP ′ a(L + )| a=1 − zP ′ a(L − )| a=1 + z∇(L + ) + z∇(L − ) = z 2 P ′ a(L 0 )| a=1 .Taking the n-th coefficient, we getp n (L + ) − p n (L − ) + c n−1 (L + ) + c n−1 (L − ) = p n−1 (L 0 ). (4.8)The skein relation for C n is obtained directly from the Conway skein relation.It depends on the number m 0 <strong>of</strong> the components in L 0 :⎧⎪⎨ C n−1 (L 0 ), if m 0 = m − 1C n (L + )−C n (L − ) =⎪⎩ 2 ∑ ∑c i (L ′ )c n−i−1 (L 0 L ′ (4.9)) if m 0 = m + 1,n−1L ′ ⊂L 0 i=0where the summation is over all sublinks L ′ <strong>of</strong> L 0 which contain both newcomponents appearing after the smoothing. Now Theorem 4.3.10 <strong>and</strong> equality(4.9) yieldAD n,m (G + ) − AD n,m (G − ) + C n (L + ) − C n (L − ) = AD n−1,m0 (G 0 ) + C n−1 (L 0 ).Deducting w(G 0 )c n−2 (L 0 ) from both sides <strong>of</strong> this equation <strong>and</strong> noticing thatw(G 0 ) = w(G + ) − 1 = w(G − ) + 1 <strong>and</strong> c n−1 (L + ) − c n−1 (L − ) = c n−2 (L 0 ), sow(G 0 )c n−2 (L 0 ) = w(G + )c n−1 (L + ) − w(G + )c n−1 (L + ) − c n−1 (L + ) − c n−1 (L − ),89


we obtain the desired skein relation for I n (L) + C n (L):(I n (L + ) + C n (L + )) − (I n (L − ) + C n (L − ))+c n−1 (L + ) + c n−1 (L − ) =I n−1 (L 0 ) + C n−1 (L 0 ).It remains <strong>to</strong> compare values <strong>of</strong> I n (L) + C n (L) <strong>and</strong> p n (L) on an m-componentunlink O m . From the definition <strong>of</strong> I n (L) we get I n (O m ) = AD n,m (O m ) = 0 forany n <strong>and</strong> m. Also, the equality p n (O m ) = C n (O m ) holds for any n <strong>and</strong> m, sincep 0 (O 2 ) = C 0 (O 2 ) = 2 <strong>and</strong> p n (O m ) = C n (O m ) = 0 otherwise. This concludes thepro<strong>of</strong> <strong>of</strong> the theorem.Example 4.3.12. Let G be a Gauss diagram <strong>of</strong> a link H 2 shown in Figure4.23. Let us calculate C n (H 2 ) <strong>and</strong> I n (H 2 ). Both components <strong>of</strong> H 2 are trivial,41 2 3++++Figure 4.23: Link H 2 <strong>and</strong> its Gauss diagramso C 0 (H 2 ) = 2 <strong>and</strong> C n (H 2 ) = 0 for n ≠ 0. The only ascending states <strong>of</strong> Gare {1, 2}, {1, 4}, <strong>and</strong> {3, 4}; the only descending state <strong>of</strong> G is {2, 3}. ThusAD 2,2 (G) = 4. Note that c 1 (H 2 ) = 2 <strong>and</strong> c n (H 2 ) = 0 for n ≠ 1, thusI 2 (H 2 ) = 4 − 4 · 2 = −4 <strong>and</strong> I n (H 2 ) = 0 for n ≠ 2. Indeed, one may check thatP (H 2 ) = a −3 z −1 − a −5 z −1 + a −3 z + a −1 z, so zP ′ a(H 2 )| a=1 = 2 − 4z 2 .4.3.3 The case <strong>of</strong> knotsIn this subsection we define for every n ≥ 0 another two <strong>invariants</strong> I A,n <strong>and</strong>I D,n <strong>of</strong> classical knots.Definition 4.3.13. Let G be a based Gauss diagram <strong>of</strong> a knot K. We go onthe circle <strong>of</strong> G starting from the base point until we return <strong>to</strong> the base point.Denote by w A (G) (respectively w D (G)) sum <strong>of</strong> signs <strong>of</strong> all arrows <strong>of</strong> G whichwe pass first at the arrowhead (respectively arrowtail).Theorem 4.3.14. Let G be any based Gauss diagram <strong>of</strong> a knot K. Then forevery n ≥ 0 bothI A,n (G) := A 2 n,1(G) − w A (G) · c n−1 (K),I D,n (G) := D 2 n,1(G) − w D (G) · c n−1 (K)90


are <strong>invariants</strong> <strong>of</strong> a knot K.These <strong>invariants</strong> will be denoted by I A,n (K) <strong>and</strong> I D,n (K) respectively.Pro<strong>of</strong>. We will prove the invariance <strong>of</strong> I A,n (G); the pro<strong>of</strong> for I D,n (G) is thesame.A well-known fact in knot theory is that for classical knots, theories <strong>of</strong> closed<strong>and</strong> long knots are equivalent. Thus it suffices <strong>to</strong> prove the invariance <strong>of</strong> I A,n (G)under Reidemeister moves Ω 1 – Ω 3 applied away from the base point. Note thatboth w A (G) <strong>and</strong> A 2 n,1(G) are invariant under Ω 2 <strong>and</strong> Ω 3 (see Lemma 4.3.3). Itremains <strong>to</strong> prove the invariance <strong>of</strong> I A,n (G) under Ω 1 .Let G <strong>and</strong> ˜G be two based Gauss diagrams that differ by an application <strong>of</strong>Ω 1 , so that ˜G has an additional isolated arrow α either as in Figure 4.24a, oras in Figure 4.24b.εεabFigure 4.24: Two versions <strong>of</strong> the Reidemeister move Ω 1In the first case, A 2 n,1( ˜G) = A 2 n,1(G) <strong>and</strong> w A ( ˜G) = w A (G), thus we haveI A,n (G) = I A,n ( ˜G).In the second case, by Corollary 4.2.5 we getA 2 n,1( ˜G) = A 2 n,1(G) + εc n−1 (K).We also have w A ( ˜G) = w A (G) + ε, <strong>and</strong> thus again I A,n (G) = I A,n ( ˜G).Note that for every G we have w(G) = w A (G) + w D (G); also, for knots onehas I n (K) = p n (K).Corollary 4.3.15. For every knot KI n (K) = I A,n (K) + I D,n (K) = p n (K).91


Bibliography[1] Afanasiev D.: On a Generalization <strong>of</strong> Alex<strong>and</strong>er Polynomial for Long Virtual<strong>Knot</strong>s, preprint arXiv:math.GT/0906.4245, 2009.[2] Alex<strong>and</strong>er J.W.: Topological <strong>invariants</strong> <strong>of</strong> knots <strong>and</strong> links, Trans. Amer.Math. Soc. 30, no. 2 (1928), 275-306.[3] Arnold V. I.: Mathematical methods <strong>of</strong> classical mechanics, translated fromthe 1974 Russian edition by K. Vogtmann <strong>and</strong> A. Weinstein, GraduateTexts in Math., Vol. 60, Springer-Verlag, New York <strong>and</strong> Berlin, 1978.[4] Baader S.: Asymp<strong>to</strong>tic Rasmussen invariant, arXiv: math.GT/0702335,2007.[5] Banyaga A.: Sur la structure du groupe des diffeomorphismes qui preserventune forme symplectique, (French) Comment. Math. Helv. 53, no. 2(1978), 174-227.[6] Bar-Natan D.: On Khovanov’s categorification <strong>of</strong> the Jones polynomial,Alg. Geom. Top., 2 (2002), 337-370.[7] Bar-Natan D.: On the Vassiliev knot <strong>invariants</strong>, Topology, 34 (1995), 423-472.[8] Barge J., Ghys E.: Cocycles d’Euler et de Maslov, Math. Ann. 294:2 (1992),235-265.[9] Bavard C.: Longueur stable des commutateurs, Enseign. Math. (2) 37, no.1-2 (1991), 109-150.[10] Beliakova A., Wehrli S.: Categorification <strong>of</strong> the colored Jones polynomial<strong>and</strong> Rasmussen invariant <strong>of</strong> links, Canad. J. Math. 60, no. 6 (2008), 1240-126692


[11] Biran P., En<strong>to</strong>v M., Polterovich L.: Calabi <strong>quasi</strong>morphisms for the symplecticball, Communications in Contemporary Mathematics, 6:5 (2004),793-802.[12] Brooks R.: Some remarks on bounded cohomology, Riemann surfaces <strong>and</strong>related <strong>to</strong>pics: Proceedings <strong>of</strong> the 1978 S<strong>to</strong>ny Brook Conference (StateUniv. New York, S<strong>to</strong>ny Brook, N.Y., 1978), Ann. <strong>of</strong> Math. Studies, 97,Prince<strong>to</strong>n Univ. Press, Prince<strong>to</strong>n, 1981, 53-63.[13] Calabi E.: On the group <strong>of</strong> au<strong>to</strong>morphisms <strong>of</strong> a symplectic manifold, Problemsin Analysis, symposium in honour <strong>of</strong> S. Bochner, ed. R.C. Gunning,Prince<strong>to</strong>n Univ. Press (1970), 1-26.[14] Calegari D.: scl, MSJ Memoirs 20, Mathematical Society <strong>of</strong> Japan (2009).[15] Chmu<strong>to</strong>v S., Cap Khoury M., Rossi A.: Polyak-Viro formulas for coefficients<strong>of</strong> the Conway polynomial, Journal <strong>of</strong> <strong>Knot</strong> Theory <strong>and</strong> Its Ramifications18, no. 6 (2009), 773-783.[16] Chmu<strong>to</strong>v S., Dushin S., Mos<strong>to</strong>voy J.: CDBook:Introduction <strong>to</strong> Vassiliev<strong>Knot</strong> Invariants, http://www.math.ohio-state.edu/∼chmu<strong>to</strong>v/preprints/.[17] Chmu<strong>to</strong>v S., Polyak M.: Elementary combina<strong>to</strong>rics for HOMFLYPT polynomial,Int. Math. Res. Notices (2009), doi:10.1093/imrn/rnp137.[18] Chow W.: On the algebraic braid group, Ann. Math., 49 (1948), 654-658.[19] Chrisman M.: On the Goussarov-Polyak-Viro finite-type <strong>invariants</strong> <strong>and</strong> thevirtualization move, arXiv: math.GT/1001.2247, 2010.[20] Conway J.: An enumeration <strong>of</strong> knots <strong>and</strong> links, Computational problemsin abstract algebra, Ed.J.Leech, Pergamon Press, (1969), 329-358.[21] De la Harpe P.: Topics in geometric group theory, University <strong>of</strong> ChicagoPress, 2000.[22] Ebin D. G., Marsden J.: Groups <strong>of</strong> diffeomorphisms <strong>and</strong> the motion <strong>of</strong> anincompressible fluid, Ann. <strong>of</strong> Math. 92, (1970), 102-163.[23] Fathi A., Laudenbach F., Poenaru V.: Travaux de Thurs<strong>to</strong>n sur les surfaces,Asterisque, 66-67. Societe Mathematique de France, Paris, (1979).93


[24] Freyd P., Yetter D., Hoste J., Lickorish W. B. R., Millett K., OcneanuA.: A new polynomial invariant <strong>of</strong> knots <strong>and</strong> links, Bull. AMS 12 (1985),239-246.[25] Gambaudo J.M., Ghys E.: Enlacements asymp<strong>to</strong>tiques, Topology 36, no.6 (1997), 1355-1379.[26] Gambaudo J.M., Ghys E.: Commuta<strong>to</strong>rs <strong>and</strong> diffeomorphisms <strong>of</strong> surfaces,Ergodic Theory Dynam. Systems 24, no. 5 (2004), 1591-1617.[27] Gambaudo J.M., Ghys E.: Braids <strong>and</strong> signatures, Bull. Soc. Math. France133, no. 4 (2005), 541-579.[28] Gambaudo J.M., Lagrange M.: Topological lower bounds on the distancebetween area preserving diffeomorphisms, Bol. Soc. Brasil. Mat. 31, (2000),1-19. CMP 2000:12.[29] Goussarov M.: On n-equivalence <strong>of</strong> knots <strong>and</strong> <strong>invariants</strong> <strong>of</strong> finite degree,Topology <strong>of</strong> manifolds <strong>and</strong> varieties, Adv. Soviet Math. 18 AMS (1994),173-192.[30] Goussarov M., Polyak M., Viro O.: Finite type <strong>invariants</strong> <strong>of</strong> classical <strong>and</strong>virtual knots, Topology 39 (2000), 1045-1068.[31] Grigorchuk R.: Some results on bounded cohomology, Combina<strong>to</strong>rial <strong>and</strong>Geometric Group Theory, LMS Lecture Notes Ser. 284, Cambridge UniversityPress, (1994), 111-163.[32] Hedden M., Ording P.: The Ozsvath-Szabo <strong>and</strong> Rasmussen concordance<strong>invariants</strong> are not equal, American Journal <strong>of</strong> Mathematics 130(2) (2008),441-453.[33] Jones V.F.R.: A polynomial invariant for knots via von Neumann algebras,Bull. Amer. Math. Soc. 12 (1985), 103-112.[34] Kauffman L.: State models <strong>and</strong> the Jones polynomial, Topology 26 (1987),no. 3, 395-407.[35] Kauffman L.: Virtual <strong>Knot</strong> Theory, European J. Comb. Vol. 20 (1999),663-690.94


[36] Khovanov M.: A categorification <strong>of</strong> the Jones polynomial, Duke Math J.,101 (2000), 359-426.[37] Khovanov M., Rozansky L.: Matrix fac<strong>to</strong>rizations <strong>and</strong> link homology,Fund. Math. 199 (1) (2008), 1-91.[38] Kishino T., Sa<strong>to</strong>h S.: A note on non-classical virtual knots, Journal <strong>of</strong><strong>Knot</strong> Theory <strong>and</strong> its Ramifications 13 no. 7 (2004), 845-856.[39] Lee E. S.: An endomorphism <strong>of</strong> the Khovanov invariant, arXiv:math.GT/0210213, 2004.[40] Lickorish W. B. R.: An Introduction <strong>to</strong> <strong>Knot</strong> Theory, 1997 Springer-VerlagNew York, Inc.[41] Lickorish W. B. R., Millett K.: A polynomial invariant <strong>of</strong> oriented links,Topology 26 (1) (1987), 107-141.[42] Livings<strong>to</strong>n C.: Computations <strong>of</strong> the Ozsvath–Szabo knot concordance invariant,Geometry <strong>and</strong> Topology 8 (2004), 735-742.[43] Manolescu C., Ozsvath P., Sarkar S: A combina<strong>to</strong>rial description <strong>of</strong> knotFloer homology, Ann. <strong>of</strong> Math. (2) 169 (2009), 633-660 MR2480614.[44] Matveev S., Polyak M.: A simple formula for the Casson-Walker invariant,Journal <strong>Knot</strong> Theory 18 (2009), 841-864.[45] McDuff D., Salamon D.: Introduction <strong>to</strong> Symplectic Topology, OxfordMathematical Monographs, Clarendon Press, 1995.[46] Milnor John W.: Lectures on the h-cobordism theorem, notes by L. Siebenmann<strong>and</strong> J. Sondow, Prince<strong>to</strong>n University Press, Prince<strong>to</strong>n, NJ, (1965).[47] Murasugi K.: <strong>Knot</strong> theory <strong>and</strong> its <strong>applications</strong>, Translated from the 1993Japanese original by Bohdan Kurpita. Birkhäuser Bos<strong>to</strong>n, Inc., Bos<strong>to</strong>n,(1996).[48] Östlund O.-P.: Invariants <strong>of</strong> knot diagrams <strong>and</strong> relations among Reidemeistermoves, Journal <strong>of</strong> <strong>Knot</strong> Theory <strong>and</strong> Its Ramifications 10, no. 8(2001), 1215-1227.95


[49] Ozsvath P. S., Szabo Z.: Holomorphic discs <strong>and</strong> knot <strong>invariants</strong>, Adv.Math., 186 (1):58-116, 2004.[50] Ozsvath P. S., Szabo Z.: <strong>Knot</strong> Floer homology <strong>and</strong> the four-ball genus,Geometry <strong>and</strong> Topology 7 (2003), 615-639.[51] Polterovich L.: The geometry <strong>of</strong> the group <strong>of</strong> symplectic diffeomorphisms,Lectures in Mathematics ETH Zurich, Birkhauser, Basel, 2001. xii+132pages.[52] Polyak M.: Minimal sets <strong>of</strong> Reidemeister moves, preprintarXiv:math.GT/0908.3127, 2009.[53] Polyak M., Viro O.: Gauss diagram formulas for Vassiliev <strong>invariants</strong>, Int.Math. Res. Notices 11 (1994), 445-454.[54] Przytycki J., Traczyk P.: Invariants <strong>of</strong> links <strong>of</strong> the Conway type, Kobe J.Math. 4 (1988), 115-139.[55] Rasmussen J. A.: Floer homology <strong>and</strong> knot complements, PhD thesis, HarvardUniversity, 2003.[56] Rasmussen J. A.: Khovanov homology <strong>and</strong> the slice genus, arXiv:math.GT/0402131, 2004.[57] Rolfsen D.: <strong>Knot</strong>s <strong>and</strong> links, Mathematics Lecture Series, No. 7. Publishor Perish, Inc., Berkeley, Calif., 1976. ix+439 pp.[58] Sawollek J.: On Alex<strong>and</strong>er-Conway polynomials for virtual knots <strong>and</strong> links,Journal <strong>of</strong> <strong>Knot</strong> Theory Ramifications 12, no. 6 (2003), 767-779.[59] Shumakovitch A.: Rasmussen invariant, slice-Bennequin inequality, <strong>and</strong>sliceness <strong>of</strong> knots, Journal <strong>of</strong> <strong>Knot</strong> Theory <strong>and</strong> its Ramifications (2007),1403-1412.[60] Silver D., Williams S.: Alex<strong>and</strong>er Groups <strong>and</strong> Virtual Links, Journal <strong>of</strong><strong>Knot</strong> Theory <strong>and</strong> Its Ramifications, 10 (2001), 151-160.[61] Silver D., Williams S.: Polynomial <strong>invariants</strong> <strong>of</strong> virtual links, Journal <strong>of</strong><strong>Knot</strong> Theory <strong>and</strong> its Ramifications 12 (2003), 987-1000.[62] Viro O.: Remarks on definition <strong>of</strong> Khovanov homology,arXiv:math.GT/0202199v1.96

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!