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

Quasimorphisms on surfaces and continuity in the Hofer norm

Michael Khanevsky Michael Khanevsky, Mathematics Department, Technion - Israel Institute of Technology Haifa, 32000, Israel khanev@math.technion.ac.il
Abstract.

There is a number of known constructions of quasimorphisms on Hamiltonian groups. We show that on surfaces many of these quasimorphisms are not compatible with the Hofer norm in a sense they are not continuous and not Lipschitz. The only exception known to the author is the Calabi quasimorphism on a sphere  [EP] and the induced quasimorphisms on genus-zero surfaces (e.g.  [BEP]).

The author was supported by the Azrieli Fellowship.

1. Introduction and results

Let (M,ω)(M,\omega) be a symplectic manifold. The Hamiltonian group of MM admits Hofer’s metric which, roughly speaking, measures mechanical energy needed to deform one Hamiltonian into another. This metric has many nice properties - it is bi-invariant and on closed manifolds the induced geometry is a natural one: any other bi-invariant CC^{\infty}-continuous Finsler metric is equivalent to Hofer’s ( [BO]). However until now the large scale geometry not well understood as there is a very limited set of tools available to estimate distances. It is particularly difficult to produce lower bounds.

Ideally, one would like to construct an invariant that estimates the distance and at the same time respects the group structure - namely, a homomorphism. In the case when the symplectic form ω\omega is exact the Calabi homomorphism is such a tool as it is 1-Lipschitz with respect to the metric. However, by a well known result  [Ban], Ham(M)Ham(M) is either simple (when MM is closed) or it contains a simple subgroup of codimension 1, therefore all homomorphisms factor through Calabi. As an attempt to relax the constraints one may consider quasimorphisms: maps Ham(M)Ham(M)\to\mathbb{R} which are additive up to a bounded defect. When one is interested in coarse estimates of geometry, the defect does not play a significant role. There is a number of known constructions of quasimorphisms for various manifolds and the main question is whether they can be utilized to extract geometric information. For example, the Calabi quasimorphism on Ham(S2)Ham(S^{2}) and certain other manifolds ( [EP]) arises as a Floer-theoretic spectral invariant and is naturally Lipschitz with respect to the Hofer metric. There is a series of other constructions that consider topological invariants of orbits of a Hamiltonian flow. Some of these quasimorphisms can be seen as generalizations of the Calabi homomorphism so there was a hope they may inherit metric properties. Unfortunately, this is not the case and all constructions known to the author except for the Calabi quasimorphisms and the induced ones are not Lipschitz with respect to Hofer’s metric. In this article we review two families of quasimorphisms on surfaces constructed by Polterovich and Gambaudo-Ghys and show that they are neither Lipschitz nor continuous in the Hofer metric. Some results can be generalized to symplectic manifolds of higher dimension.

The situation is different when one considers some other (not bi-invariant) metrics on Ham(M)Ham(M). For example, the quasimorphisms mentioned above are continuous in the LpL^{p} norm (see  [BS3, BS1, BS2]).

Acknowledgements: The author is grateful to L. Buhovski, L. Polterovich and E. Shelukhin for useful discussions and comments.

2. Preliminaries

Let (M,ω)(M,\omega) be a symplectic manifold, gg a Hamiltonian diffeomorphism with compact support in MM. The Hofer norm g\|g\| (see  [Hof]) is defined by

g=infG01max(G(,t)minG(,t))dt\|g\|=\inf_{G}\int_{0}^{1}\max\left(G(\cdot,t)-\min G(\cdot,t)\right)\mathrm{d}t

where the infimum goes over all compactly supported Hamiltonian functions G:M×[0,1]G:M\times[0,1]\to\mathbb{R} such that gg is the time-1 map of the induced flow. The Hofer metric is given by

dH(g1,g2)=g1g21.d_{H}(g_{1},g_{2})=\|g_{1}g_{2}^{-1}\|.

Let GG be a group. A function r:Gr:G\to\mathbb{R} is called a quasimorphism if there exists a constant DD (called the defect of rr) such that |r(fg)r(f)r(g)|<D|r(fg)-r(f)-r(g)|<D for all f,gGf,g\in G. The quasimorphism rr is called homogeneous if it satisfies r(gm)=mr(g)r(g^{m})=mr(g) for all gGg\in G and mm\in\mathbb{Z}. Any homogeneous quasimorphism satisfies r(fg)=r(f)+r(g)r(fg)=r(f)+r(g) for commuting elements f,gf,g. Every quasimorphism is equivalent (up to a bounded deformation) to a unique homogeneous one  [Cal]. A quasimorphism is called genuine if it is not a homomorphism.

Lemma 1.

Let GG be a Lie group equipped with a bi-invariant path metric, r:Gr:G\to\mathbb{R} a homogeneous quasimorphism. Then rr is Lipschitz if and only if it is continuous at the identity.

Proof.

The fact that Lipschitz property implies continuity is obvious. We show the opposite direction. Suppose rr is continuous at the identity, namely, given ε>0\varepsilon>0, there exists δ=δ(ε)>0\delta=\delta(\varepsilon)>0 such that any gGg\in G with dist(g,𝟙)=gδ\operatorname{dist}(g,\mathbb{1})=\|g\|\leq\delta satisfies |r(g)|<ε|r(g)|<\varepsilon. Fix an ε>0\varepsilon>0 and select an appropriate δ=δ(ε)\delta=\delta(\varepsilon).

The Lipschitz property holds on a large scale: let g,hGg,h\in G with gg arbitrary and hδ\|h\|\geq\delta. By the triangle inequality,

|r(hg)r(g)||r(h)|+D|r(hg)-r(g)|\leq|r(h)|+D

where DD denotes the defect of rr.

Pick a path hth_{t} connecting hh to the identity with len(ht)h+δ\operatorname{len}(h_{t})\leq\|h\|+\delta. The path hth_{t} can be cut into N:=len(ht)δ<hδ+2N:=\left\lceil\frac{\operatorname{len}(h_{t})}{\delta}\right\rceil<\frac{\|h\|}{\delta}+2 arcs of length at most δ\delta, hence hh can be presented as a composition of NN elements in the δ\delta-neighborhood of the identity. Therefore

|r(h)|=|r(h1hN)|<N(ε+D)<(hδ+2)(ε+D)h3δ(ε+D).|r(h)|=|r(h_{1}\cdot\ldots\cdot h_{N})|<N\cdot\left(\varepsilon+D\right)<\left(\frac{\|h\|}{\delta}+2\right)\left(\varepsilon+D\right)\leq\|h\|\cdot\frac{3}{\delta}(\varepsilon+D).

Finally,

|r(hg)r(g)||r(h)|+D<h3δ(ε+D)+hDδ|r(hg)-r(g)|\leq|r(h)|+D<\|h\|\cdot\frac{3}{\delta}(\varepsilon+D)+\|h\|\cdot\frac{D}{\delta}

and the property holds since h=dist(h,𝟙)=dist(hg,g)\|h\|=\operatorname{dist}(h,\mathbb{1})=\operatorname{dist}(hg,g) by bi-invariance of the metric.

We show that the Lipschitz property holds also locally: pick g,hGg,h\in G with gg arbitrary and hh in the δ\delta-neighborhood of the identity. Denote

k:=δh>δ2h.k:=\left\lfloor\frac{\delta}{\|h\|}\right\rfloor>\frac{\delta}{2\|h\|}.

Then by homogenuity and the triangle inequality

|r(hg)r(g)|\displaystyle|r(hg)-r(g)| =\displaystyle= 1k|r((hg)k)r(gk)|=\displaystyle\frac{1}{k}\cdot\left|r\left((hg)^{k}\right)-r\left(g^{k}\right)\right|=
1k|r(hghg1g2hg2gk1hg1kgk)r(gk)|\displaystyle\frac{1}{k}\cdot\left|r\left(h\cdot ghg^{-1}\cdot g^{2}hg^{-2}\cdot\ldots\cdot g^{k-1}hg^{1-k}\cdot g^{k}\right)-r\left(g^{k}\right)\right|\leq
1k(|r(hghg1g2hg2gk1hg1k)|+D)\displaystyle\frac{1}{k}\cdot\left(\left|r\left(h\cdot ghg^{-1}\cdot g^{2}hg^{-2}\cdot\ldots\cdot g^{k-1}hg^{1-k}\right)\right|+D\right)

By bi-invariance gnhgn=hδk\|g^{n}hg^{-n}\|=\|h\|\leq\frac{\delta}{k} and the triangle inequality implies

hghg1g2hg2gk1hg1kδ.\left\|h\cdot ghg^{-1}\cdot g^{2}hg^{-2}\cdot\ldots\cdot g^{k-1}hg^{1-k}\right\|\leq\delta.

Therefore using our choice of δ\delta,

|r(hg)r(g)|<1k(ε+D)<2hδ(ε+D)=2δ(ε+D)dist(hg,g).|r(hg)-r(g)|<\frac{1}{k}\cdot(\varepsilon+D)<\frac{2\|h\|}{\delta}(\varepsilon+D)=\frac{2}{\delta}(\varepsilon+D)\cdot\operatorname{dist}(hg,g).

Combining the results above, rr is globally Lipschitz. ∎

Therefore a non-Lipschitz quasimorphism is also non-continuous. (With some extra effort one may show it is nowhere continuous.) In what follows we will concentrate only on the Lipschitz property.

3. Calabi quasimorphisms

Let Ft:MF_{t}:M\to\mathbb{R}, t[0,1]t\in[0,1] be a time-dependent smooth function with compact support. Define

Cal~(Ft)=01(MFtω)dt.\widetilde{Cal}(F_{t})=\int_{0}^{1}\left(\int_{M}F_{t}\omega\right)\mathrm{d}t.

When ω\omega is exact on MM, Cal~\widetilde{Cal} descends to a homomorphism Cal:Ham(M)Cal:Ham(M)\to\mathbb{R} which is called the Calabi homomorphism. For any gHam(M)g\in Ham(M), Cal(g)gCal(g)\leq\|g\| by its definition.

Equip the unit sphere S2S^{2} with a symplectic form normalized by S2ω=1\int_{S^{2}}\omega=1.  [EP] presents construction of a homogeneous quasimorphism CalS2:Ham(S2)Cal_{S^{2}}:Ham(S^{2})\to\mathbb{R} based on a Floer-theoretical spectral invariant of Ham(S2)Ham(S^{2}). Being a spectral invariant, it is Lipschitz with respect to the Hofer norm.

A symplectic embedding j:MS2j:M\to S^{2} of a genus zero surface into a sphere induces the pullback quasimorphism j(CalS2)j^{*}(Cal_{S^{2}}) on Ham(M)Ham(M). In  [BEP] the authors show that this construction provides continuum of linearly independent quasimorphisms, all Lipschitz in the Hofer norm. These quasimorphisms provide useful tools to analyze the Hofer geometry of Ham(M)Ham(M). For example, existence of such quasimorphisms allows to construct quasi-isometric embeddings of “large” sets into Ham(M)Ham(M) (see  [EP]) or into spaces of Hamiltonian-isotopic Lagrangian submanifolds of MM (e.g.  [Kha, Sey]).

These quasimorphisms can be seen as generalizations of the Calabi homomorphism as they coincide with the Calabi invariant given a Hamiltonian isotopy which is supported in a displaceable subset. There was a hope that other quasimorphisms (some of them can also be seen as certain generalizations of Calabi) may share geometric properties with the Calabi homomorphism, for example, turn to be Lipschitz. Unfortunately, it is not the case. We analyze two constructions of quasimorphisms in the sections below.

4. Polterovich construction

4.1. The quasimorphisms

Let (M,ω)(M,\omega) be a symplectic manifold with non-abelian fundamental group. We equip it with an auxiliary Riemannian metric and pick a basepoint zMz\in M. Given a non-trivial homogeneous quasimorphism r:π1(M,z)r:\pi_{1}(M,z)\to\mathbb{R} one constructs a quasimorphism ρ:Ham(M)\rho:Ham(M)\to\mathbb{R} as follows. For each xMx\in M choose a short geodesic path γx\gamma_{x} which connects xx to zz. Given a Hamiltonian isotopy hth_{t} between 𝟙\mathbb{1} and hHam(M)h\in Ham(M), for each xMx\in M denote by lx(ht)=γx{ht(x)}γh(x)l_{x}(h_{t})=\gamma_{x}*\{h_{t}(x)\}*-\gamma_{h(x)} the closed loop defined by concatenation of geodesic paths with the trajectory of xx under hth_{t}. Let

ρ^(ht)=Mr([lx(ht)])ω\hat{\rho}(h_{t})=\int_{M}r([l_{x}(h_{t})])\omega

where [lx(ht)]π1(M,z)[l_{x}(h_{t})]\in\pi_{1}(M,z) is the equivalence class of lx(ht)l_{x}(h_{t}). ρ^:Ham(M)~\hat{\rho}:\widetilde{Ham(M)}\to\mathbb{R} is a quasimorphism and its homogenization

ρ(ht)=limk1kMr([lx((ht)k)])ω\rho(h_{t})=\lim_{k\to\infty}\frac{1}{k}\int_{M}r([l_{x}((h_{t})^{k})])\omega

does not depend on the choices of metric, zz, geodesic paths or hth_{t}, resulting in a homogeneous quasimorphism ρ:Ham(M)\rho:Ham(M)\to\mathbb{R}. For more details see  [Pol].

4.2. Noncontinuity

We start with a simple example which will be extended to the general case later. Let MM be a closed surface of genus two equipped with a symplectic form ω\omega. Denote by a,b,a,ba,b,a^{\prime},b^{\prime} the standard generators of π1(M,)\pi_{1}(M,*) and pick a quasimorphism r:π1(M,)r:\pi_{1}(M,*)\to\mathbb{R} such that r(a)=r(b)=0,r(aba1b1)=1r(a)=r(b)=0,r(aba^{-1}b^{-1})=1 (for example, a word counting quasimorphism). Let ρ:Ham(M)\rho:Ham(M)\to\mathbb{R} be the quasimorphism induced by rr. Denote by PP a pair of pants whose boundary components represent the free homotopy classes of aba1b1,a,a1aba^{-1}b^{-1},a,a^{-1}.

Refer to caption
Figure 1. Pair of pants

We construct a Hamiltonian HH from the indicator function of PP which is cut off near the boundary. Note that the flow ϕHt\phi_{H}^{t} generated by HH is controlled by the cutoff and is supported in a tubular neighborhood of the boundary P\partial P. For a reasonable choice of the cutoff (when the level sets {H1(c)}c(0,1)\{H^{-1}(c)\}_{c\in(0,1)} have exactly one connected component near each connected component of P\partial P) support of the flow consists of three strips around P\partial P. These strips are foliated by periodic trajectories in the free homotopy classes of a,a1,aba1b1a,a^{-1},aba^{-1}b^{-1}. The classes a,a1a,a^{-1} do not affect the value of ρ(ϕHt)\rho(\phi_{H}^{t}) and contribution comes only from trajectories in the class aba1b1aba^{-1}b^{-1}. A simple computation shows that ρ(ϕHt)=t\rho(\phi_{H}^{t})=t and it is independent of the cutoff. (Intuitively, when one applies a steeper cutoff, velocity of the flow increases linearly with the slope and inversely proportional to its area of support, so faster rotation is compensated by the reduced volume of motion.)

We adjust the cutoff so that nearly all non-stationary points near P\partial P have a fixed rational period. Pick an area preserving parametrization S1×(ε,ε)S^{1}\times(-\varepsilon,\varepsilon) with coordinates (θ,h)(\theta,h) near each connected component of P\partial P (in our notation S1=/S^{1}=\mathbb{R}/\mathbb{Z}). Suppose that PP is defined in these coordinates by {h0}\{h\geq 0\}. Pick a natural number n>1/εn>1/\varepsilon and a smoothing parameter δ<<1/n\delta<<1/n, define the cutoff function c:(ε,ε)[0,1]c:(-\varepsilon,\varepsilon)\to[0,1] by

c(h)={0if h0hnif δh1nδ1if h1nc(h)=\left\{\begin{array}[]{ll}0&\textrm{if $h\leq 0$}\\ hn&\textrm{if $\delta\leq h\leq\frac{1}{n}-\delta$}\\ 1&\textrm{if $h\geq\frac{1}{n}$}\end{array}\right.

and interpolate it smoothly in the intervals {0hδ},{1nδh1n}\left\{0\leq h\leq\delta\right\},\left\{\frac{1}{n}-\delta\leq h\leq\frac{1}{n}\right\}. Let H(θ,h)=c(h)H(\theta,h)=c(h). As the result, all points in the annulus {δh1nδ}\left\{\delta\leq h\leq\frac{1}{n}-\delta\right\} have rational period T=1/nT=1/n and the remaining non-stationary points in PP are contained in a region of area 6δ6\delta and have controlled dynamics in a sense that their period is at least 1/n1/n. Note that the smoothing parameter δ\delta can be chosen arbitrarily small and can be reduced independently of PP and nn. Such adjustment of δ\delta does not increase the maximal velocity of the flow.

Step I: Given N>>1N>>1 consider a pair of pants PP as above with Area(P)<<1N2\operatorname{Area}(P)<<\frac{1}{N^{2}}. We place NN deformed copies P1=h1(P),,PN=hN(P)P_{1}=h_{1}(P),\ldots,P_{N}=h_{N}(P) of PP (where the deformations hih_{i} are symplectic near PP and isotopic to the identity) in a way so that no triple intersections of PiP_{i} occur.

Refer to caption
Figure 2.

Pick a Hamiltonian function HH supported in PP as described above. The cutoff of HH is adjusted in a way that most (up to a region of area 6δ<<Area(P)6\delta<<\operatorname{Area}(P)) non-stationary points have the same period T=1/nT=1/n for some nn\in\mathbb{N}. Denote the resulting Hamiltonian by HPH_{P} and let HiH_{i} be the deformations of HPH_{P} supported in each PiP_{i} defined by hi(Hi)=HPh_{i}^{*}(H_{i})=H_{P}. Let ϕit=hi,ϕHPt\phi_{i}^{t}=h_{i,*}\phi^{t}_{H_{P}} denote their Hamiltonian flows.

Step II: Pick a system of disjoint open neighborhoods {Uij}1i<jN\left\{U_{ij}\right\}_{1\leq i<j\leq N} such that PiPjUijP_{i}\cap P_{j}\subset U_{ij}. For i>ji>j let Uij=UjiU_{ij}=U_{ji}. We assume that in each connected component QQ of UijU_{ij} there is enough “empty space” not occupied by PiP_{i} and PjP_{j}: Area(Q)>3Area((PiPj)Q)\operatorname{Area}(Q)>3\operatorname{Area}((P_{i}\cup P_{j})\cap Q). Otherwise we may either repeat the construction applying the same deformations hih_{i} to a narrower pair of pants PP or deform the symplectic form ω\omega in the complement of the union of all pairs of pants and redistribute area to satisfy this condition.

Let t0t_{0} be the minimal time required for any of the flows ϕi\phi_{i} to travel between two neighborhoods UklU_{kl} and UklU_{k^{\prime}l^{\prime}}. Namely, given t<t0t<t_{0}, ϕit(Ukl)Ukl=\phi_{i}^{t}(U_{kl})\cap U_{k^{\prime}l^{\prime}}=\emptyset for all {k,l}{k,l}\{k,l\}\neq\{k^{\prime},l^{\prime}\} and all 1iN1\leq i\leq N. Pick a natural mm such that τ:=Tm=1mn<t02\tau:=\frac{T}{m}=\frac{1}{mn}<\frac{t_{0}}{2}.

Step III: We make a series of adjustments to this construction. The goal is to ensure that most points are TT-periodic while the rest (that are beyond our direct control) will not travel too far, hence their contribution to the quasimorphism will be small.

First of all, we modify the smoothing parameter δ\delta that appears in the construction of HPH_{P}. Fix an auxiliary Riemannian metric on MM. Let vv be the maximal point velocity of the flows ϕi\phi_{i} and dQd_{Q} be the maximal diameter of connected components QQ of the neighborhoods UijU_{ij}. Let

Sτ={[γ]π1(M,)|len(γ)τv+dQ+2diam(M)}S_{\tau}=\{[\gamma]\in\pi_{1}(M,*)\,\big{|}\,\operatorname{len}(\gamma)\leq\tau v+d_{Q}+2\operatorname{diam}(M)\}

be a set of homotopy classes of short loops and rτ=max{r(c)|cSτ}r_{\tau}=\max\{r(c)\,\big{|}\,c\in S_{\tau}\} be the maximal value of the quasimorphism r:π1(M,)r:\pi_{1}(M,*)\to\mathbb{R} restricted to SτS_{\tau}. (Short loops represent a finite number of homotopy classes so the maximum is attained.) We pick

δ<min(τ20(rτ+D)N,τm3N)\delta<\min\left(\frac{\tau}{20(r_{\tau}+D)N},\;\frac{\tau m}{3N}\right)

where DD is the defect of rr. HiH_{i} and the deformed flows ϕi\phi_{i} are adjusted according to this value of the smoothing parameter δ\delta. Note that this modification does not increase the maximal velocity of the flow. Hence τ\tau is still less than the travel time between different neighborhoods UklU_{kl}.

Refer to caption
Figure 3. Local adjustments

Furthermore, perturb P1,,PNP_{1},\ldots,P_{N} inside the neighborhoods UijU_{ij} so that

Area(PiPj)δmN.\operatorname{Area}(P_{i}\cap P_{j})\leq\frac{\delta}{mN}.

We apply the same perturbations to the Hamiltonian functions HiH_{i} and the flows ϕi\phi_{i}. This modification does not affect dynamics of the flows outside the neighborhoods UijU_{ij}, hence the minimal travel time between neighborhoods is still greater than τ\tau. Moreover, the modified flows may have increased velocities locally inside UijU_{ij}, but on a larger scale (time τ\tau and above) they travel approximately the same distance.

Let Φt=ϕ1tϕNt\Phi^{t}=\phi_{1}^{t}\circ\ldots\circ\phi_{N}^{t} be the composition of the time-tt maps of the flows. We claim that for t=τt=\tau,

ρ(Φτ)=Nτ+O(δ)>(N2)τ\rho(\Phi^{\tau})=N\tau+O(\delta)>(N-2)\tau

while the Hofer norm is bounded by 2τ2\tau. Taking NN\to\infty, we deduce that ρ\rho is not Lipschitz.

Refer to caption
Figure 4. The composition

By the composition formula, the flow Φt\Phi^{t} is generated by the time-dependent Hamiltonian

G(t)=H1+(ϕ1t)(H2)++(ϕ1tϕN1t)(HN)G(t)=H_{1}+\left(\phi_{1}^{-t}\right)^{*}(H_{2})+\ldots+\left(\phi_{1}^{-t}\circ\ldots\circ\phi_{N-1}^{-t}\right)^{*}(H_{N})

Namely, the pairs of pants PiP_{i} which support the Hamiltonian functions HiH_{i} are deformed and “dragged” along the flows. For time t<t0t<t_{0} supports of the summands cannot have triple intersections: that happens only after some PiP_{i} is dragged by ϕjt\phi_{j}^{t} to a point where it intersects some (possibly deformed) PkP_{k}. That means that certain points pUijp\in U_{ij} have arrived in (or passed through) other neighborhood UjkU_{jk} and that cannot happen before t=t0t=t_{0}. As each summand admits values in the interval [0,1][0,1], maxGtminGt2\max G_{t}-\min G_{t}\leq 2 for all 0t<t00\leq t<t_{0} and the bound on the Hofer norm follows.

We estimate the value of the quasimorphism ρ(Φτ)\rho(\Phi^{\tau}). Intuitively speaking, the time-τ\tau maps {ϕiτ}\{\phi_{i}^{\tau}\} “almost commute” (commute in the complement of a subset of area 7Nδ7N\delta), hence their contribution to the quasimorphism “almost adds up” (up to a defect which is small compared to the values ρ(ϕiτ)\rho(\phi_{i}^{\tau}) and is controlled by δ\delta).

Pick kNk\leq N. We consider three ϕkτ\phi_{k}^{\tau}-invariant regions in MM according to the dynamics of points under ϕkτ\phi_{k}^{\tau}:

  1. (1)

    points outside PkP_{k} and those points inside that are stationary under ϕk\phi_{k}. They do not contribute to ρ(ϕkτ)\rho(\phi_{k}^{\tau}).

  2. (2)

    the set AkA_{k} of mm-periodic points under ϕkτ\phi_{k}^{\tau} that never hit any of the remaining pairs of pants {Pi}ik\{P_{i}\}_{i\neq k}.

    All non-stationary points are contained in the three cutoff strips near Pk\partial P_{k}. Each strip has area 1n\frac{1}{n} and we have to remove those points that are not mm-periodic due to smoothing. Thus the set of mm-periodic points has area bounded between 3n6δ\frac{3}{n}-6\delta and 3n\frac{3}{n}. After removing points whose trajectories intersect {Pi}ik\{P_{i}\}_{i\neq k} we get the estimate

    Area(Ak)3n6δmikArea(PiPk)3n6δmNδmN=3n7δ\operatorname{Area}(A_{k})\geq\frac{3}{n}-6\delta-m\sum_{i\neq k}\operatorname{Area}(P_{i}\cap P_{k})\geq\frac{3}{n}-6\delta-mN\cdot\frac{\delta}{mN}=\frac{3}{n}-7\delta

    Points of AkA_{k} are distributed in three strips near Pk\partial P_{k} and their trajectories, depending on the strip, represent either a,a1a,a^{-1} or aba1b1aba^{-1}b^{-1}. aa and a1a^{-1} do not affect ρ(ϕkτ)\rho(\phi_{k}^{\tau}) while the set AkA^{\prime}_{k} of points in the class aba1b1aba^{-1}b^{-1} has Area(Ak)1n3δ\operatorname{Area}(A^{\prime}_{k})\geq\frac{1}{n}-3\delta and contributes

    Area(Ak)r(aba1b1)m1nm3δm=τ3δm\frac{\operatorname{Area}(A^{\prime}_{k})\cdot r(aba^{-1}b^{-1})}{m}\geq\frac{1}{nm}-\frac{3\delta}{m}=\tau-\frac{3\delta}{m}
  3. (3)

    the set BkB_{k} containing the rest of points (non-stationary points whose period is different from mm or whose trajectory under ϕkτ\phi_{k}^{\tau} hits other pairs of pants). By a similar computation, Area(Bk)7δ\operatorname{Area}(B_{k})\leq 7\delta. While we have less control on the dynamics of BkB_{k}, we note that the velocity of points in BkB_{k} under the flow ϕk\phi_{k} is bounded by vv outside neighborhoods QQ of the intersections PkPjP_{k}\cap P_{j} (it can still be very large inside QQ because of the adjustments). Also note that the time needed to travel between different neighborhoods is larger than τ\tau. Hence for each pBkp\in B_{k}, the trajectory {ϕkt(p))}t[0,τ]\{\phi_{k}^{t}(p))\}_{t\in[0,\tau]} has length at most τv+dQ\tau v+d_{Q}.

    Applying stabilization, the trajectory of pp under (ϕkτ)K(\phi_{k}^{\tau})^{K} has length bounded by K(τv+dQ)K(\tau v+d_{Q}) and [lp((ϕkτ)K)]\left[l_{p}\left((\phi_{k}^{\tau})^{K}\right)\right] can be presented as a product of at most KK elements from SτS_{\tau}, so r([lp((ϕkτ)K)])K(rτ+D)r([l_{p}((\phi_{k}^{\tau})^{K})])\leq K(r_{\tau}+D). The contribution of BkB_{k} to ρ(ϕkτ)\rho(\phi_{k}^{\tau}) is bounded in the absolute value by

    limK1KBk|r([lp((ϕkτ)K)])|ω(rτ+D)Area(Bk)=7δ(rτ+D).\lim_{K\to\infty}\frac{1}{K}\int_{B_{k}}|r([l_{p}((\phi_{k}^{\tau})^{K})])|\omega\leq(r_{\tau}+D)\cdot\operatorname{Area}(B_{k})=7\delta(r_{\tau}+D).

Now note that for all kNk\leq N, AkA_{k} is Φτ\Phi^{\tau}-invariant and the action of Φτ\Phi^{\tau} coincides with that of ϕkτ\phi_{k}^{\tau}. Indeed, for a point pAkp\in A_{k}, (Φτ)K(p)=(ϕkτ)K(p)(\Phi^{\tau})^{K}(p)=(\phi_{k}^{\tau})^{K}(p) for all KK as pp never hits the pairs of pants {Pi}ik\{P_{i}\}_{i\neq k}. More than that, the trajectory {Φt(p)}t[0,τ]\{\Phi^{t}(p)\}_{t\in[0,\tau]} is homotopic relative endpoints to {ϕkt(p)}t[0,τ]\{\phi_{k}^{t}(p)\}_{t\in[0,\tau]} (other ingredients of Φt(p)=ϕ1tϕNt(p)\Phi^{t}(p)=\phi_{1}^{t}\circ\ldots\circ\phi_{N}^{t}(p) can be homotoped away). The same is true for iterations of Φτ\Phi^{\tau}, hence AkA_{k} contributes to ρ(Φτ)\rho(\Phi^{\tau}) the same amount as to ρ(ϕkτ)\rho(\phi_{k}^{\tau}). We finish with an observation that the sets {Ak}k=1N\{A_{k}\}_{k=1}^{N} are disjoint and do not intersect B=j=1NBjB=\bigcup_{j=1}^{N}B_{j}.

Let pBp\in B. If pPNp\in P_{N}, the trajectory {Φt(p)}t[0,τ]\{\Phi^{t}(p)\}_{t\in[0,\tau]} is the trajectory {ϕNt(p)}t[0,τ]\{\phi_{N}^{t}(p)\}_{t\in[0,\tau]} which was possibly deformed by the flows ϕN1,,ϕ1\phi_{N-1},\ldots,\phi_{1}. Such deformation occurs each time {ϕNt(p)}t[0,τ]\{\phi_{N}^{t}(p)\}_{t\in[0,\tau]} crosses a pair of pants PN1,,P1P_{N-1},\ldots,P_{1}. Within time τ\tau only one such crossing is possible. As the result, {Φt(p)}t[0,τ]\{\Phi^{t}(p)\}_{t\in[0,\tau]} is homotopic (relative endpoints) to a path of length at most (2τv+dQ)(2\tau v+d_{Q}) (the path can be longer than τv+dQ\tau v+d_{Q} when {ϕNt(p)}t[0,τ]\{\phi_{N}^{t}(p)\}_{t\in[0,\tau]} either starts or ends at the intersection with other pair of pants). The same argument can be applied to points pB(PN1PN)p\in B\cap(P_{N-1}\setminus P_{N}), pB(PN2(PNPN1))p\in B\cap(P_{N-2}\setminus(P_{N}\cup P_{N-1})) and so on by induction. After taking KK iterations, trajectory of any pBp\in B is homotopic relative endpoints to a path shorter than 2K(τv+dQ)2K(\tau v+d_{Q}), thus the contribution of BB to ρ(Φτ)\rho(\Phi^{\tau}) in the absolute value is at most

limK1KB|r([lp((Φτ)K)])|ω\displaystyle\lim_{K\to\infty}\frac{1}{K}\int_{B}|r([l_{p}((\Phi^{\tau})^{K})])|\omega \displaystyle\leq 2(rτ+D)Area(B)2(rτ+D)i=1NArea(Bi)\displaystyle 2(r_{\tau}+D)\cdot\operatorname{Area}(B)\leq 2(r_{\tau}+D)\cdot\sum_{i=1}^{N}\operatorname{Area}(B_{i})
\displaystyle\leq 14Nδ(rτ+D).\displaystyle 14N\delta(r_{\tau}+D).

As a summary, points from A=i=1NAiA=\bigcup_{i=1}^{N}A_{i} contribute to ρ(Φτ)\rho(\Phi^{\tau}) at least Nτ3NδmN\tau-\frac{3N\delta}{m}. B=i=1NBiB=\bigcup_{i=1}^{N}B_{i} alters the result by at most 14Nδ(rτ+D)14N\delta(r_{\tau}+D). Points pM(AB)p\in M\setminus(A\cup B) are stationary and do not contribute. We deduce

ρ(Φτ)Nτ3Nδm14Nδ(rτ+D)>(N2)τ.\rho(\Phi^{\tau})\geq N\tau-\frac{3N\delta}{m}-14N\delta(r_{\tau}+D)>(N-2)\tau.

4.3. General case

Let MM be a symplectic surface of finite type and suppose r:π1(M,)r:\pi_{1}(M,*)\to\mathbb{R} is a genuine homogeneous quasimorphism. Pick a,bπ1(M,)a,b\in\pi_{1}(M,*) such that r(ab)r(a)+r(b)r(ab)\neq r(a)+r(b) (without loss of generality assume that dr=r(a)+r(b)r(ab)>0d_{r}=r(a)+r(b)-r(ab)>0). Let α,β\alpha,\beta be based loops representing a,ba,b, respectively. Denote by ρ\rho be the induced quasimorphism on Ham(M)Ham(M). Let γ=αβ\gamma=\alpha*\beta. We wish to construct a Hamiltonian diffeomorphism Φτ\Phi^{\tau} such that ρ(Φτ)>>ΦτH\rho(\Phi^{\tau})>>\|\Phi^{\tau}\|_{H}.

Step I: Pick a small tubular neighborhood PP of γ\gamma. Applying C0C^{0}-small perturbations to α,β,γ\alpha,\beta,\gamma we may assume that the three curves intersect (and self-intersect) transversely in a finite number of points and still lie in PP.

Step II: Consider NN deformed copies (N>>1N>>1) P1=h1(P),,PN=hN(P)P_{1}=h_{1}(P),\ldots,P_{N}=h_{N}(P) of PP where the deformations hih_{i} are isotopic to the identity and area-preserving near PP so that:

  • no triple intersections of {Pi}i=1N\{P_{i}\}_{i=1}^{N} occur

  • the collection of loops Q={αi=hi(α)}i{βi=hi(β)}i{γi=hi(γ)}iQ=\{\alpha_{i}=h_{i}(\alpha)\}_{i}\cup\{\beta_{i}=h_{i}(\beta)\}_{i}\cup\{\gamma_{i}=h_{i}(\gamma)\}_{i} does not have triple intersection/self-intersection points, the total number of intersections is finite and they are all transverse.

Step III: Construct a time-dependent Hamiltonian flow Φt\Phi^{t} supported in a narrow tubular neighborhood of curves from QQ which translates points along {αi}i=1N,{βi}i=1N\{\alpha_{i}\}_{i=1}^{N},\{\beta_{i}\}_{i=1}^{N} preserving orientation of curves and along {γi}i=1N\{\gamma_{i}\}_{i=1}^{N} with reversed orientation.

Refer to caption
Figure 5. Local picture

Local picture:

  • near a curve cQc\in Q and away from intersection points the flow is a parallel translation of a narrow strip around cc with fixed velocity and flux 11. The flow is cut off in a smooth way so that it remains parallel to cc and the area affected by the cutoff is small.

  • near an intersection of two curves ci,cjc_{i},c_{j} (or a self-intersection of two arcs ci,cjc_{i},c_{j} of the same curve from QQ): construct the flows gi,gjg_{i},g_{j} parallel to ci,cjc_{i},c_{j} as in the previous case. We may symplectically perturb gi,gjg_{i},g_{j} near cicjc_{i}\cap c_{j} to ensure that Area(supp(gi)supp(gj))\operatorname{Area}(\operatorname{supp}(g_{i})\cap\operatorname{supp}(g_{j})) is sufficiently small. Define the local flow Φt\Phi^{t} to be the composition gitgjtg_{i}^{t}\circ g_{j}^{t}. Intuitively speaking, the curve cjc_{j} and the flow gjg_{j} following cjc_{j} “drift” with gitg_{i}^{t}.

For tt small enough the flow charts can be adjusted and patched together to form a global symplectic time-dependent flow Φt\Phi^{t}. More than that, using local adjustments, one can ensure that:

  • flow strips around αi,βi,γi\alpha_{i},\beta_{i},\gamma_{i} are narrow enough and stay inside PiP_{i}.

  • local velocities are chosen in a way so that if one ignores all intersections, time TT needed to traverse a loop cQc\in Q is the same for all loops in QQ.

  • the total area of the cutoff zones and intersections / self-intersections of the flow strips is at most δ\delta (δ\delta is a small parameter which will be specified later).

The construction described above works for tt0t\leq t_{0} where t0t_{0} is the minimal time needed for an intersection point pcicjp\in c_{i}\cap c_{j} to leave the appropriate chart when it follows the flows gig_{i} or gjg_{j}.

Clearly, the flow Φt\Phi^{t} is area-preserving. We claim that it has zero flux, hence is Hamiltonian. Indeed, the flux along each αi,βi\alpha_{i},\beta_{i} is 11 while the flux along γi\gamma_{i} is 1-1 due to reversed orientation. A generic cycle cc in (M,M)(M,\partial M) satisfies #cαi+#cβi=#cγi\#c\cap\alpha_{i}+\#c\cap\beta_{i}=\#c\cap\gamma_{i} (the intersection points are counted with signs), so the fluxes cancel out. Denote by HtH_{t} the Hamiltonian function which generates Φt\Phi^{t}. It is a well known fact that Ht(p)Ht(q)H_{t}(p)-H_{t}(q) equals the flux of Φt\Phi^{t} through a curve which connects pp to qq (for a Hamiltonian flow the flux depends on the endpoints and not on the curve itself). The argument above also shows that the flux between two points p,qPip,q\notin\bigcup P_{i} is zero, hence, up to a shift by appropriate constant, suppHtPi\operatorname{supp}H_{t}\subseteq\bigcup P_{i}.

For a generic curve cc connecting two points p,qP1(α1β1γ1)p,q\in P_{1}\setminus(\alpha_{1}\cup\beta_{1}\cup\gamma_{1}) let Kc=#cα1+#cβ1#cγ1K_{c}=\#c\cap\alpha_{1}+\#c\cap\beta_{1}-\#c\cap\gamma_{1} (count with signs). Roughly, if one ignores the width of flow strips around the curves, KcK_{c} measures the flux between pp and qq of the flows along α1,β1,γ1\alpha_{1},\beta_{1},\gamma_{1}. K=maxcKcK=\max_{c}K_{c} equals the maximal flux between any two points in P1P_{1}. Fluxes of flows inside each PiP_{i} are additive, so as long as no triple intersections are allowed and tt0t\leq t_{0} (meaning the curves do not drift far and combinatorics of their intersections does not change), the maximal flux of Φt\Phi^{t} (and hence the variation of HtH_{t}) is at most 2K2K. That is, the Hofer length of the isotopy Φt\Phi^{t} is bounded by 2Kt2Kt. Note that the bound KK may depend on the perturbations performed in Step I but does not depend on the choice of NN in Step II.

Pick τ=Tmt0\tau=\frac{T}{m}\leq t_{0} for some mm\in\mathbb{N}. The flow strip around a curve cQc\in Q has flux 11 and period TT, hence has area equal to TT (up to effects of smoothing which slow down the flow and are controlled by δ\delta). We subtract points which are slowed by smoothing and also those whose trajectory under {(Φτ)k}k\{(\Phi^{\tau})^{k}\}_{k\in\mathbb{Z}} lands on the intersections/self-intersections of strips. The remaining region of area TO(δ)T-O(\delta) is mm-periodic and contributes to ρ(Φτ)\rho(\Phi^{\tau}) either

Tr(a)m+O(δ)=τr(a)+O(δ)T\cdot\frac{r(a)}{m}+O(\delta)=\tau r(a)+O(\delta)

in the case c=αic=\alpha_{i}, τr(b)+O(δ)\tau r(b)+O(\delta) if c=βic=\beta_{i} or τr(ab)+O(δ)-\tau r(ab)+O(\delta) when c=γic=\gamma_{i} for some iNi\leq N.

The rest of non-stationary points (either affected by smoothing or those visiting intersections of strips) are contained in a region BB of area O(δ)O(\delta). The long-term dynamics of pBp\in B can be complicated, but locally (till time τ\tau) pp can visit at most one intersection of strips, so the length of trajectory {Φt(p)}t[0,τ]\{\Phi^{t}(p)\}_{t\in[0,\tau]} is uniformly bounded. As the result, |1kr([lp((Φτ)k)])|\left|\frac{1}{k}r([l_{p}((\Phi^{\tau})^{k})])\right| is also uniformly bounded by some C>0C>0 and the contribution of BB to ρ(Φτ)\rho(\Phi^{\tau}) is bounded in the absolute value by

Area(B)C=O(δ).\operatorname{Area}(B)\cdot C=O(\delta).

Summing up,

ρ(Φτ)=i=1N(τr(a)+τr(b)τr(ab)+O(δ))+O(δ)=τNdr+O(δ).\rho(\Phi^{\tau})=\sum_{i=1}^{N}(\tau r(a)+\tau r(b)-\tau r(ab)+O(\delta))+O(\delta)=\tau Nd_{r}+O(\delta).

Now we pick δ=δ(N)\delta=\delta(N) small enough to ensure ρ(Φτ)>τ(N1)dr\rho(\Phi^{\tau})>\tau(N-1)d_{r} and perform rearrangements of smoothing and intersection zones to comply with this chosen value of δ\delta. These local adjustments do not affect t0t_{0} or estimates of trajectory lengths, hence all computations we performed above remain in place. The claim follows by picking NN such that Ndr>>2KNd_{r}>>2K.

4.4. Higher dimensions

The construction above can be lifted to a symplectic ball bundle M^M\widehat{M}\to M. We lift the flow Φt\Phi^{t} from MM and cut it off near M^\partial\widehat{M}. When the cutoff is steep enough, it affects only a small volume of motion while keeping dynamics under control (cutoff keeps the MM component of the flow velocity bounded and introduces a fast rotation around the boundary in the fiber coordinates. This rotation does not affect average homotopy classes of orbits.) Hence its effect on the quasimorphism ρ^:Ham(M^)\widehat{\rho}:Ham(\widehat{M})\to\mathbb{R} can be made arbitrarily small. The bounds on the Hofer length in M^\widehat{M} remain the same as in MM and non-continuity follows.

Let MM be a higher dimension symplectic manifold, r:π(M,)r:\pi(M,*)\to\mathbb{R} a genuine homogeneous quasimorphism. Let α,β\alpha,\beta be based loops in MM such that r([αβ])r([α])+r([β])r([\alpha*\beta])\neq r([\alpha])+r([\beta]). Perturb α,β\alpha,\beta so that they intersect at one point and that near the intersection they lie in a symplectic 22-disk DD. DD can be extended to a smooth symplectic surface Σ\Sigma with boundary which contains both α\alpha and β\beta. By the symplectic neighborhood theorem, a small neighborhood of Σ\Sigma in MM is symplectomorphic to a ball bundle over Σ\Sigma. Now we may apply the construction for ball bundles and show non-continuity of the quasimorphism induced by rr.

Remark 2.

Lev Buhovsky has pointed out that in the case of ball bundles there is a much simpler construction of Hamiltonian flows which also shows non-continuity of quasimorphisms ( [Buh]). However this construction does not apply in the two-dimensional case so we could not avoid the more complicated setup.

5. Gambaudo-Ghys construction

5.1. The quasimorphisms

Let (M,ω)(M,\omega) be a symplectic surface of finite type and area 11, equipped with an auxiliary Riemannian metric. Denote by BkB_{k} the braid group with kk strands in MM and by PBkPB_{k} the pure braid group. Suppose {gt}Ham(M)\{g_{t}\}\in Ham(M) is an isotopy starting at the identity and ending at gg. Pick kk distinct points w1,,wkw_{1},\ldots,w_{k} in MM to be a basepoint 𝐰\mathbf{w} in the configuration space Xk(M)X_{k}(M). For an 𝐱=(x1,,xk)Xk(M)\mathbf{x}=(x_{1},\ldots,x_{k})\in X_{k}(M) we construct kk based loops lwj,xj(gt)l_{w_{j},x_{j}}(g_{t}) by concatenating a short geodesic path γwj,xj\gamma_{w_{j},x_{j}} with the trajectory {gt(xj)}\{g_{t}(x_{j})\} and closing the loop by a short geodesic path γg(xj),wj\gamma_{g(x_{j}),w_{j}}. For almost all 𝐱Xk\mathbf{x}\in X_{k}, the kk-tuple of loops l𝐱(gt)=(lw1,x1(gt),,lwk,xk(gt))l_{\mathbf{x}}(g_{t})=\left(l_{w_{1},x_{1}}(g_{t}),\ldots,l_{w_{k},x_{k}}(g_{t})\right) defines a based loop in Xk(M)X_{k}(M). π1(Xk,𝐰)\pi_{1}(X_{k},\mathbf{w}) can be identified with the pure braid group PBkPB_{k}, hence [l𝐱(gt)][l_{\mathbf{x}}(g_{t})] defines an element in PBkPB_{k}.

Let r:Bkr:B_{k}\to\mathbb{R} be a homogeneous quasimorphism. Then

ρ^(gt)=Xkr([l𝐱(gt)])ωk\hat{\rho}(g_{t})=\int_{X_{k}}r\left([l_{\mathbf{x}}(g_{t})]\right)\omega^{k}

defines a quasimorphism ρ^:Ham(M)~\hat{\rho}:\widetilde{Ham(M)}\to\mathbb{R} and its homogenization

ρ(gt)=limm1mXkr([l𝐱((gt)m)])ωk\rho(g_{t})=\lim_{m\to\infty}\frac{1}{m}\int_{X_{k}}r\left([l_{\mathbf{x}}((g_{t})^{m})]\right)\omega^{k}

does not depend on the choices of metric, basepoint 𝐰\mathbf{w}, geodesic paths or gtg_{t}, resulting in a homogeneous quasimorphism ρ:Ham(M)\rho:Ham(M)\to\mathbb{R}. For a more detailed explanation see  [Bra].

This construction generalizes the quasimorphisms by Polterovich: PB1(M)π1(M,)PB_{1}(M)\simeq\pi_{1}(M,*), therefore for k=1k=1 both constructions agree.

Remark 3.

In the case of a disk M=D2M=D^{2}, PB2(M)PB_{2}(M)\simeq\mathbb{Z}. In this case any homogeneous quasimorphism r:PB2(M)r:PB_{2}(M)\to\mathbb{R} is a homomorphism and the induced quasimorphism ρ:Ham(D2)\rho:Ham(D^{2})\to\mathbb{R} is a multiple of the Calabi homomorphism (see  [GtG, Fat]).

This way, the Gambaudo-Ghys construction can be seen as a generalization of the Calabi homomorphism.

5.2. Noncontinuity

Denote by FkF_{k} the subgroup of PBkPB_{k} defined by fixing strands 22 to kk and letting the first strand wind around. FkF_{k} fits into a short exact sequence

0FkPBkPBk100\to F_{k}\to PB_{k}\to PB_{k-1}\to 0

where the map PBkPBk1PB_{k}\to PB_{k-1} is the Chow homomorphism which “forgets” the first strand of BkB_{k}.

We consider a particular family of quasimorphisms: suppose r:Bkr:B_{k}\to\mathbb{R} is a homogeneous quasimorphism such that the restriction r|Fkr\big{|}_{F_{k}} is a genuine quasimorphism. We show that the quasimorphism ρ:Ham(M)\rho:Ham(M)\to\mathbb{R} induced by the Gambaudo-Ghys construction is not Hofer-Lipschitz.

Remark 4.

E. Shelukhin pointed out that in the case of surfaces with boundary, PBkPB_{k} may be presented as a semidirect product of FkF_{k} with PBk1PB_{k-1} (PBk1PB_{k-1} is embedded into PBkPB_{k} by adding a stationary strand near the boundary). In this case one can show by induction that every genuine homogeneous quasimorphism on BkB_{k} restricts to a genuine quasimorphism on FkF_{k}.

If MM has no boundary, consider a punctured surface M˙=M{p}\dot{M}=M\setminus\{p\}. The natural inclusion M˙M\dot{M}\to M induces a surjective homomorphism j:Bk(M˙)Bk(M)j:B_{k}(\dot{M})\to B_{k}(M) of braid groups. In this case the pullback quasimorphism jr:Bk(M˙)j^{*}r:B_{k}(\dot{M})\to\mathbb{R} satisfies the condition by the remark above, therefore its restriction jr|Fk(M˙)j^{*}r\big{|}_{F_{k}(\dot{M})} is a genuine quasimorphism. But

jr|Fk(M˙)=j|Fk(M˙)(r|Fk(M)),j^{*}r\big{|}_{F_{k}(\dot{M})}=j\big{|}_{F_{k}(\dot{M})}^{*}\left(r\big{|}_{F_{k}(M)}\right),

hence r|Fk(M)r\big{|}_{F_{k}(M)} is genuine as well.

Therefore this technical condition on r|Fkr\big{|}_{F_{k}} is automatically satisfied.

When M=S2M=S^{2} and k3k\leq 3, BkB_{k} is finite and all homogeneous quasimorphisms are trivial. Otherwise FkF_{k} is naturally isomorphic to the fundamental group of k1k-1 times punctured surface M=M{p2,,pk}M^{\prime}=M\setminus\{p_{2},\ldots,p_{k}\} (see  [Bir]). Let r:π1(M,)r^{\prime}:\pi_{1}(M^{\prime},*)\to\mathbb{R} be the quasimorphism induced by r|Fkr\big{|}_{F_{k}}. Pick a,bπ1(M,)a,b\in\pi_{1}(M^{\prime},*) such that r(a)+r(b)r(ab)r^{\prime}(a)+r^{\prime}(b)\neq r^{\prime}(ab).

We use construction from the previous section and apply a variation of the argument by T. Ishida  [Ish]. Given N>>1N>>1 and sufficiently small δ=δ(N)\delta=\delta(N), consider a narrow region PNP_{N} which contains loops representing classes a,b,aba,b,ab. We place NN copies of PNP_{N} in MM^{\prime} without triple intersections and construct ΦNτNHam(M)\Phi_{N}^{\tau_{N}}\in Ham(M^{\prime}). Let gN=(ΦNτN)mNg_{N}=\left(\Phi_{N}^{\tau_{N}}\right)^{m_{N}} where mNm_{N} is the common period for most non-stationary points under ΦNτN\Phi_{N}^{\tau_{N}}.

gNg_{N} has several invariant subsets:

  • subsets Aa,Ab,AabA_{a},A_{b},A_{ab} that are fixed by gNg_{N}. Trajectories of points in AaA_{a} represent aa, while AbA_{b} and AabA_{ab} represent bb and (ab)-(ab), respectively. Each of the three subsets has area NTN+O(δ)=NmNτN+O(δ)NT_{N}+O(\delta)=Nm_{N}\tau_{N}+O(\delta),

  • subset BB with Area(B)=O(δ)\operatorname{Area}(B)=O(\delta). Trajectories in BB are beyond control and have topological complexity of O(mN)O(m_{N}),

  • all remaining points in MM^{\prime} are stationary under the flow.

In addition, the Hofer norm gNM2mNτN=2TN\|g_{N}\|_{M^{\prime}}\leq 2m_{N}\tau_{N}=2T_{N} and δ\delta can be chosen arbitrarily small without affecting the estimates above.

Pick ε<<1\varepsilon<<1 and d2,dk[0,1]d_{2},\ldots d_{k}\in[0,1] such that ε+d2++dk=1\varepsilon+d_{2}+\ldots+d_{k}=1. Rescale the symplectic form of MM^{\prime} so that Area(M)=ε\operatorname{Area}(M^{\prime})=\varepsilon and glue in disks D2,,DkD_{2},\ldots,D_{k} of area d2,,dkd_{2},\ldots,d_{k} in the punctures p2,,pkp_{2},\ldots,p_{k}, respectively. The resulting surface M^\widehat{M} has area 11 hence is symplectomorphic to the unpunctured surface MM. gNg_{N} induce Hamiltonian diffeomorphisms on M^\widehat{M} which will be denoted by gNg_{N} as well, and due to rescaling their Hofer norms satisfy

gNM^2TNε.\|g_{N}\|_{\widehat{M}}\leq 2T_{N}\varepsilon.

We estimate ρ(gN)\rho(g_{N}). Consider possible braids that arise from trajectories of 𝐱=(x1,xk)Xk(M)\mathbf{x}=(x_{1},\ldots x_{k})\in X_{k}(M) by gNg_{N} and their contribution to ρ(gN)\rho(g_{N}):

  • all kk points belong to the disks {Dj}\{D_{j}\}, hence are stationary. Such braids are trivial hence do not affect the value of ρ\rho.

  • one point lies in MM^{\prime} and exactly one point belongs to each disk DjD_{j}, 2jn2\leq j\leq n. Order of points does not matter: reordering of points corresponds to permutation of strands. rr is homogeneous, hence invariant under conjugations in BkB_{k} that permute stands. Without loss of generality, assume x1Mx_{1}\in M^{\prime}. l𝐱(gN)l_{\mathbf{x}}(g_{N}) is induced by the trajectory of x1x_{1} in MM^{\prime} and k1k-1 stationary strands in the disks.

    r([l𝐱(gN)])=r([lx1(gN)])r([l_{\mathbf{x}}(g_{N})])=r^{\prime}([l_{x_{1}}(g_{N})])

    (the same is true for iterations of gNg_{N}). AaAbAabA_{a}\cup A_{b}\cup A_{ab} contribute

    (NTN+O(δ))(r(a)+r(b)r(ab))εd2dk\left(NT_{N}+O(\delta)\right)\left(r^{\prime}(a)+r^{\prime}(b)-r^{\prime}(ab)\right)\varepsilon d_{2}\ldots d_{k}

    while BB contributes

    O(δ)O(mN)εd2dk=O(δ)εd2dkO(\delta)\cdot O\left(m_{N}\right)\varepsilon d_{2}\ldots d_{k}=O(\delta)\varepsilon d_{2}\ldots d_{k}

    as δ\delta can be chosen independently of mNm_{N}. Stationary points in MM^{\prime} induce a trivial braid and do not contribute anything. The total contribution is

    (k!NTN(r(a)+r(b)r(ab))+O(δ))εd2dk=\left(k!NT_{N}(r^{\prime}(a)+r^{\prime}(b)-r^{\prime}(ab))+O(\delta)\right)\varepsilon d_{2}\ldots d_{k}=
    =(NTNR1,2,,dk+O(δ))εd2dk.=\left(NT_{N}R_{1,2,\ldots,d_{k}}+O(\delta)\right)\varepsilon d_{2}\ldots d_{k}.

    k!k! takes into account reorderings of points and the resulting coefficient

    R1,2,,dk0R_{1,2,\ldots,d_{k}}\neq 0

    .

  • xiMx_{i}\in M^{\prime} for some 1ik1\leq i\leq k while the rest belong to disks {Dj}\{D_{j}\} and the correspondence between points and disks is not bijective. Let i2,,iki_{2},\ldots,i_{k} denote the indices of disks for the points {xj}ji\{x_{j}\}_{j\neq i} in the ascending order of jj. The contribution of this configuration of points is given by

    RNTN,i,i2,,ikεdi2dikR_{NT_{N},i,i_{2},\ldots,i_{k}}\varepsilon d_{i_{2}}\ldots d_{i_{k}}

    for some coefficient RNTN,i,i2,,ikR_{NT_{N},i,i_{2},\ldots,i_{k}}\in\mathbb{R}. Similarly to the previous case, the dynamics of gNg_{N} implies that this coefficient scales linearly with NTNNT_{N} (up to O(δ)O(\delta)), so we may rewrite

    RNTN,i,i2,,ikεdi2dik=(NTNRi,i2,,ik+O(δ))εdi2dik.R_{NT_{N},i,i_{2},\ldots,i_{k}}\cdot\varepsilon d_{i_{2}}\ldots d_{i_{k}}=(NT_{N}R_{i,i_{2},\ldots,i_{k}}+O(\delta))\varepsilon d_{i_{2}}\ldots d_{i_{k}}.

    Summing over all configurations, we get the following polynomial in ε,d2,,dk\varepsilon,d_{2},\ldots,d_{k}

    1ik2i2,,ikk(NTNRi,i2,,ik+O(δ))εdi2dik\sum_{\begin{subarray}{c}1\leq i\leq k\\ 2\leq i_{2},\ldots,i_{k}\leq k\end{subarray}}(NT_{N}R_{i,i_{2},\ldots,i_{k}}+O(\delta))\varepsilon d_{i_{2}}\ldots d_{i_{k}}

    which does not contain the monomial εd2dk\varepsilon d_{2}\ldots d_{k}.

  • two points or more belong to MM^{\prime}, the rest lie in disks {Dj}\{D_{j}\}. Topological complexity of the strand is O(mN)O\left(m_{N}\right) (this bound is independent of δ\delta) and the contribution is O(mNε2)O\left(m_{N}\cdot\varepsilon^{2}\right) (areas did_{i} are bounded by 11 and can be incorporated into O(mNε2)O\left(m_{N}\cdot\varepsilon^{2}\right)).

So the total value of the quasimorphism ρ(gN)\rho(g_{N}) is

NTNε[(R1,2,,dk+O(δ)NTN)d2dk+(Ri,i2,,ik+O(δ)NTN)di2dik]+O(mNε2)NT_{N}\varepsilon\left[\left(R_{1,2,\ldots,d_{k}}+\frac{O(\delta)}{NT_{N}}\right)d_{2}\ldots d_{k}+\sum\left(R_{i,i_{2},\ldots,i_{k}}+\frac{O(\delta)}{NT_{N}}\right)d_{i_{2}}\ldots d_{i_{k}}\right]+O\left(m_{N}\cdot\varepsilon^{2}\right)

The polynomial R1,2,,dkd2dk+Ri,i2,,ikdi2dikR_{1,2,\ldots,d_{k}}d_{2}\ldots d_{k}+\sum R_{i,i_{2},\ldots,i_{k}}d_{i_{2}}\ldots d_{i_{k}} is not zero, so one may find values for d2,,dkd_{2},\ldots,d_{k} in the simplex {di0,di<1}\{d_{i}\geq 0,\;\sum d_{i}<1\} where the polynomial does not vanish. Pick N>>1N>>1. δ\delta and ε\varepsilon are independent parameters. As soon as we fix NN (hence also TNT_{N} and mNm_{N}), we let δ,ε0\delta,\varepsilon\to 0 while d2,,dkd_{2},\ldots,d_{k} are rescaled keeping their ratios constant and the total area

Area(M^)=ε+d2++dk=1.\operatorname{Area}\left(\widehat{M}\right)=\varepsilon+d_{2}+\ldots+d_{k}=1.

In the limit ρ(gN)\rho(g_{N}) is proportional to NTNεNT_{N}\varepsilon while Hofer’s norm is bounded by 2TNε2T_{N}\varepsilon. It follows that ρ\rho cannot be Lipschitz.

References

  • [Ban] A. Banyaga. Sur la structure du groupe des diffeomorphismes qui preservent une forme symplectique. Comm.Math.Hel, 53:174–227, 1978.
  • [BEP] P. Biran, M. Entov, and L. Polterovich. Calabi quasimorphisms for the symplectic ball. Commun. Contemp. Math., 6(5):793–802, 2004.
  • [Bir] Joan S. Birman. Braids, Links, and Mapping Class Groups. (AM-82). Princeton University Press, 1974.
  • [BO] L. Buhovsky and Y. Ostrover. On the uniqueness of Hofer’s geometry. Geom. Funct. Anal., 21(6):1296–1330, 2011.
  • [Bra] M. Brandenbursky. Bi-invariant metrics and quasi-morphisms on groups of Hamiltonian diffeomorphisms of surfaces. Int. J. of Math., 26(9), 2013.
  • [BS1] Michael Brandenbursky and Egor Shelukhin. On the large-scale geometry of the LpL^{p}-metric on the symplectomorphism group of the two-sphere. arXiv e-prints, page arXiv:1304.7037, Apr 2013.
  • [BS2] Michael Brandenbursky and Egor Shelukhin. On the LpL^{p}-geometry of autonomous Hamiltonian diffeomorphisms of surfaces. Mathematical Research Letters, 22, 05 2014.
  • [BS3] Michael Brandenbursky and Egor Shelukhin. The LpL^{p}-diameter of the group of area-preserving diffeomorphisms of S2S^{2}. Geom. Topol., 21(6):3785–3810, 2017.
  • [Buh] L. Buhovsky. Private communication, December 2018.
  • [Cal] Danny Calegari. scl., volume 20. Tokyo: Mathematical Society of Japan, 2009.
  • [EP] M. Entov and L. Polterovich. Calabi quasimorphism and quantum homology. Int. Math. Res. Not., 2003(30):1635–1676, 2003.
  • [Fat] Albert Fathi. Transformations et homeomorphismes preservant la mesure : systemes dynamiques minimaux. PhD thesis, Universite Paris-Sud, 1980.
  • [GtG] Jean-Marc Gambaudo and Étetienne Ghys. Enlacements asymptotiques. Topology, 36(6):1355 – 1379, 1997.
  • [Hof] H. Hofer. On the topological properties of symplectic maps. Proc. Royal Soc. Edinburgh, 115(1-2):25–38, 1990.
  • [Ish] Tomohiko Ishida. Quasi-morphisms on the group of area-preserving diffeomorphisms of the 2-disk via braid groups. Proceedings of the American Mathematical Society, Series B, 1, 04 2012.
  • [Kha] M. Khanevsky. Hofer’s metric on the space of diameters. JTA, 1(4):407–416, 2009.
  • [Pol] L. Polterovich. Floer homology, dynamics and groups. In Morse Theoretic Methods in Nonlinear Analysis and in Symplectic Topology, volume 217 of NATO Sci. Ser. II: Math, Phys and Chem, pages 417–438. Springer Netherlands, 2006.
  • [Sey] Sobhan Seyfaddini. Unboundedness of the Lagrangian Hofer distance in the Euclidean ball. Electronic Research Announcements in Mathematical Sciences [electronic only], 21:1–7, 10 2013.