--- title: "Chapter 9" --- Tags: #geomtop/symplectic-topology #projects/my-talks Refs: [Projects/0000 Reading Notes/Audin-Damian Morse Theory and Floer Homology Reading Notes](Projects/0000%20Reading%20Notes/Audin-Damian%20Morse%20Theory%20and%20Floer%20Homology%20Reading%20Notes.md) # Background, Notation, Setup **Goals** :::{.theorem title="Arnold Conjecture (Symplectic Morse Inequalities?)"} Let $(W, \omega)$ be a compact symplectic manifold and \[ H: W\to \RR \] a time-dependent Hamiltonian with nondegenerate 1-periodic solutions. Then \[ \size \ts{\text{1-Periodic trajectories of }X_H} \geq \sum_{k\in \ZZ} \dim_{?} HM_k(W; \, \ZZ/2\ZZ) .\] > Here $HM_*(W)$ is the Morse homology, and *nondegenerate* means the differential of the flow at time 1 has no fixed vectors. ::: **Important Ideas for This Chapter**: :::{.theorem title="Use Broken Trajectories to Compactify"} $\mathcal{L}(x, y)$ is compact, where the compactification is given by adding in \[ \bd \mathcal{L}(x, y) = \ts{\text{"Broken Trajectories"}} \] ::: :::{.theorem title="Gluing Yields a Chain Complex"} \[ \del^2 = 0 \] ::: \newpage **Strategy**: > In the background, have a Hamiltonian $H: W\to \RR$. > Basic idea: cook up a gradient flow. 1. Define the action functional $\mathcal{A}_H$ > On an infinite-dimensional space, critical points are periodic solutions of $H$ 2. Construct the chain complex (graded vector space) $CF_*$. > Uses analog of the *index* of a critical point. 3. Define the vector field $X_H$ using $-\grad \mathcal{A}_H$. > This will be used to define $\del$ later. 4. Count the trajectories of $X_H$ 5. Show finite-energy trajectories connect critical points of $\mathcal{A}_H$. 6. Show *Gromov Compactness* for space of trajectories of finite energy 7. Define $\del$ > Uses another compactness property 8. Show space of trajectories is a manifold, plus analog of "Smale property" 9. \textbf{Show that} $\del^2 = 0$ using a gluing property 10. Show that $HF_*$ doesn't depend on $\mathcal{A}_H$ or $X_H$ 11. Show $HF_* \cong HM_*$, and compare dimensions of the vector spaces $CM_*$ and $CF_*$. \newpage **Ingredients**: - $(W, \omega, J)$ with $\omega \in \Omega^2(W)$ is a symplectic manifold - With $J: T_p W \to T_p W$ an almost complex structure, so $J^2 = -\id$. - $H \in C^\infty(W; \RR)$ a Hamiltonian - $X_H$ the corresponding symplectic gradient. - Defined by how it acts on tangent vectors in $T_x M$: \[ \omega_x(\wait , X_H(x)) = (dH)_x(\wait) .\] - Zeros of vector field $X_H$ correspond to critical points of $H$: \[ X_H(x) = 0 \iff (dH)_x = 0 .\] - Take the associated flow, assumed 1-periodic: \[ \psi^t \in C^\infty(W, W) \qquad \psi^1 = \id ,\] - Critical points of $H$ are periodic trajectories. - $u \in C^\infty(\RR\cross S^1; W)$ is a solution to the Floer equation. - The Floer equation and its linearization: \[ \mcf(u) &= \dd{u}{s} + J \dd{u}{t} + \grad_u(H) = 0 \\ \qty{d\mcf}_u(Y) &= \dd{Y}{s} + J_0 \dd{Y}{t} + S\cdot Y \\ \\ \\ &Y\in u^* TW,~ S \in C^\infty(\RR\cross S^1; \endo(\RR^{2n})) .\] \newpage - $\mcl W$ is the *free loop space* on $W$, i.e. space of contractible loops on $W$, i.e. $C^\infty(S^1; W)$ with the $C^\infty$ topology - Elements $x\in \mcl W$ can be viewed as maps $S^1\to W$. - Can extend to maps from a closed disc, $u: \bar \DD^2 \to M$. - Loops in $\mcl W$ can be viewed as maps $S^2\to W$, since they're maps $I\cross S^1\to W$ with the boundaries pinched:  - The action functional is given by \[ \mca_H: \mcl W &\to \RR\\ x &\mapsto -\int_{\DD} u^* \omega + \int_0^1 H_t(x(t)) ~dt \] - Example: $W = \RR^{2n} \implies A_H(x) = \int_0^1 \qty{H_t ~dt - p~dq}$. - A correspondence \[ \correspond{\text{Solutions to the } \\\text{Floer equation}} \iff \correspond{\text{Trajectories} \\ \text{of } \grad \mathcal{A}_H} .\] - $x, y$ periodic orbits of $H$ (nondegenerate, contractible), equivalently critical points of $\mathcal{A}_H$. \newpage - Assumption of *symplectic asphericity*, i.e. the symplectic form is zero on spheres. Statement: for every $u\in C^\infty(S^2, W)$, \[ \int_{S^2} u^* \omega = 0 \qtext{or equivalently} \inner{\omega}{\pi_2 W} = 0 .\] - Assumption of *symplectic trivialization*: for every $u\in C^\infty(S^2; M)$ there exists a symplectic trivialization of the fiber bundle $u^* TM$, equivalently \[ \inner{c_1 TW}{\pi_2 W} = 0 .\] > Locally a product of base and fiber, transition functions are symplectomorphisms. - Maslov index: used the fact that - Every path in $\gamma: I\to \Sp(2n, \RR)$ can be assigned an integer coming from a map $\tilde \gamma: I \to S^1$ and taking (approximately) its winding number. - $\mathcal{M}(x, y)$, the moduli space of contractible finite-energy solutions to the Floer equation connecting $x, y$. - After perturbing $H$ to get transversality, get a manifold - Dimension: \[ \dim \mcm(x, y) = \mu(x) - \mu(y) .\] - How we did it: - Describe as zeros of a section of a vector bundle over $\mathcal{P}^{1, p}(x, y)$ (Banach manifold modeled on the Sobolev spaces $W^{1, p}$), - Apply Sard-Smale to show $\mathcal{M}(x, y)$ is the inverse image of a regular value of some map. - Needed tangent maps to be Fredholm operators, proved in Ch. 8 and used to show transversality. - Showed $(d\mathcal{F})_u$ is a Fredholm operator of index $\mu(x) - \mu(y)$. $\qed$ \newpage # Reminder of Goals **Overall Goal**: :::{.theorem title="Symplectic Morse Inequalities"} \[ \size \ts{\text{1-Periodic trajectories of }X_H} \geq \sum_{k\in \ZZ} HM_k(W; \, \ZZ/2\ZZ) .\] ::: --- **Important Ideas for This Chapter**: :::{.theorem title="Using Broken Trajectories to Compactify"} $\mathcal{L}(x, y)$ is compact, \[ \bd \mathcal{L}(x, y) = \ts{\text{"Broken Trajectories"}} \] ::: :::{.theorem title="Using Gluing to Get a Chain Complex"} \[ \del^2 = 0 \] ::: \newpage # 9.1 and Review - Defined moduli space of (parameterized) **solutions**: \[ \mathcal{M}(x, y) &= \ts{\text{Contractible finite-energy solutions connecting }x, y } \\ \\ \mcm &= \ts{\text{\textbf{All} contractible finite-energy solutions to the Floer equation}} \\ &= \bigcup_{x, y} \mathcal{M}(x, y) .\] - The moduli space of (unparameterized) **trajectories** connecting $x, y$: \[ \mathcal{L}(x, y) \da \mathcal{M}(x, y) / \RR .\] - Use the quotient topology, define sequentially: \[ \tilde u_n \converges{n\to\infty}\too \tilde u \quad\iff\quad \exists \ts{s_n}\subset \RR \text{ such that } u_n(s_n + s, \wait) \converges{n\to\infty}\too u(s, \wait) .\] - When $\abs{\mu(x) - \mu(y)} = 1$, get a compact 0-manifold, so the number of trajectories $$n(x, y) \da \size \mathcal{L}(x, y)$$ is well-defined. - $C_k(H) \da \ZZ/2\ZZ[\ts{\text{Periodic orbits of } X_H \text{ of Maslov index } k }]$. - Finitely many since they are nondegeneracy implies they are isolated. :::{.remark} Some notation: \begin{center} \begin{tikzcd} \RR \ar[r] & \mathcal{M}(x, z)\ar[d, "\pi"] \\ & \mathcal{L}(x, z) \\ \end{tikzcd} \end{center} Hats will generally denote maps induced on quotient. ::: - Defined a differential \[ \del: C_k(H) &\to C_{k-1}(H) \\ x &\mapsto \sum_{\mu(y) = k-1} n(x, y) y \\ \\ n(x, y) &\da \size \ts{\text{Trajectories of } \grad \mathcal{A}_H \text{ connecting } x, y} \mod 2 \\ &= \size \mathcal{L}(x, y) \mod 2 .\] - Examined $\del^2$: \[ \del^2: C_{k}(H) &\to C_{k-2}(H) \\ x &\mapsto \del(\del(x)) \\ &= \del \qty{\sum_{\mu(y) = \mu(x)-1} n(x, y) y} \\ \\ &= \sum_{\mu(y) = \mu(x) - 1} n(x, y) \del(y) \\ \\ &= \sum_{\mu(y) = \mu(x) - 1} n(x, y) \qty{\sum_{\mu(z) = \mu(y)-1} n(y, z) z} \\ \\ &= \sum_{\mu(y) = \mu(x) - 1} \,\,\sum_{\mu(z) = \mu(y)-1} n(x, y) n(y, z) \,z \\ &= \sum_{\mu(z) = \mu(y) - 1} \qty{\sum_{\mu(y) = \mu(x)-1} n(x, y) n(y, z)}\,z \hspace{4em} \text{(finite sums, swap order)} ,\] so it suffices to show \[ \sum_{\mu(y) = \mu(x)-1} n(x, y) n(y, z) = 0 \qtext{when} \mu(z) = \mu(x) - 2 .\] > Easier to examine parity, so we'll show it's zero mod 2. \newpage - When $\mu(z) = \mu(x) - 2$, $\mathcal{L}(x, z)$ is a non-compact 1-manifold, so we compactify by adding in *broken trajectories* to get $\bar{\mathcal{L}}(x, y)$. - We'll then have \[ \bar{\mathcal{L}}(x, z) = \mathcal{L}(x, z) \union \bd \bar{\mathcal{L}}(x, z), \qquad \bd \bar{L}(x, z) = \bigcup_{\mu(y) = \mu(x) - 1} \mathcal{L}(x, y) \cross \mathcal{L}(y, z) ,\] which "space-ifies" the equation we want. - We'll show $\bd \bar{\mathcal{L}}(x, z)$ is a 1-manifold, which must have an even number of points, and thus \[ \sum_{\mu(y) = \mu(x)-1} n(x, y) n(y, z) = \size \qty{\bd \bar{\mathcal{L}}(x, z)} \equiv 0 \mod 2 .\] > Image here of relations between spaces! $\qed$ \newpage # Three Important Theorems ## First Theorem: Convergence to Broken Trajectories - Recall: *broken trajectories* are unions of intermediate trajectories connecting intermediate critical points. - Shown last time: a sequence of trajectories can converge to a broken trajectory, i.e. there are broken trajectories in the closure of $\mathcal{L}(x, z)$. - This theorem describes their behavior: :::{.theorem title="9.1.7: Convergence to Broken Trajectories"} Let $\ts{u_n}$ be a sequence in $\mathcal{M}(x, z)$, then there exist \ - A subsequence $\ts{u_{n_j}}$ \ - Critical points $\ts{x_0, x_1, \cdots, x_{\ell+1}}$ with $x_0=x$ and $x_{\ell+1} = z$ \ - Sequences $\ts{s_n^1}, \ts{s_n^2}, \cdots, \ts{s_n^\ell}$. \ - Elements $u^k \in \mathcal{M}(x_k, x_{k+1})$ such that for every $0\leq k \leq \ell$, \[ u_{n_j} \cdot s_n^k \converges{n\to\infty}\too u^k .\] ::: - Upshots: - Every sequence upstairs has a subsequence which (after reparameterizing) converges - This descends to actual convergence after quotienting by $\RR$? - Yields uniqueness of limits in $\mathcal{L}(x, z)$, thus a separated topology - Sequentially compact $\iff$ compact since $\mathcal{L}(x, z)$ is a metric space? :::{.corollary title="Compactness"} $\bar{\mathcal{L}}(x, z)$ is compact. ::: \newpage ## Second Theorem: Compactness of $\bar{\mathcal{L}}(x, z)$ :::{.definition title="Regular Pair"} For an almost complex structure $J$ and a Hamiltonian $H$, the pair $(H, J)$ is **regular** if the Floer map $\mathcal{F}$ is transverse to the zero section in the following vector bundle: \ \begin{center} \begin{tikzcd} E_u \da \ts{\text{Vector fields tangent to $M$ along $u$}} \ar[r] & C^\infty(\RR\cross S^1; TM)\ar[dd] \\ & \\ & C^\infty(\RR\cross S^1; M) \ar[uu, bend left, "\mathcal{F}", dotted] \ar[uu, bend right, "\mathbf{0}"', dotted]\\ \end{tikzcd} \end{center} ::: Most of chapter 9 is spent proving this theorem: :::{.theorem title="9.2.1"} Let $(H, J)$ be a regular pair with $H$ nondegenerate and $x, z$ be two periodic trajectories of $H$ such that \[ \mu(x) = \mu(z) + 2 .\] Then $\bar{\mathcal{L}}(x, z)$ is a compact 1-manifold with boundary with \[ \bd \bar{\mathcal{L}}(x, z) = \bigcup_{y\in \mathcal{I}(x, z)} \mathcal{L}(x ,y) \cross \mathcal{L}(y, z) \\ \text{where}\qquad \mathcal{I}(x, z) = \ts{ y \st \mu(x) < \mu(y) < \mu(z) } .\] > Note: possibly a typo in the book? Has $x, y$ on the LHS. ::: :::{.corollary} \[\bd^2 = 0.\] ::: \newpage ## Third Theorem: Gluing :::{.theorem title="9.2.3: Gluing"} Let $x,y,z$ be three critical points of $\mathcal{A}_H$ with three consecutive indices \[ \mu(x) = \mu(y)+1 = \mu(z) + 2 .\] and let \[ (u, v) \in \mathcal{M}(x, y) \cross \mathcal{M}(y, z) \quad\leadsto\quad (\hat u, \hat v)\in \mathcal{L}(x, y) \cross \mathcal{L}(y, z) .\] Then 1. There exists a $\rho_0 > 0$ and a differentiable map \[ \psi: [\rho_0, \infty) &\to \mathcal{M}(x, z) \] such that $\hat \psi$, the induced map on the quotient \begin{center} \begin{tikzcd} {[\rho_0, \infty)} \ar[r, "\psi"] \ar[rd, "\hat\psi"', dotted] & \mathcal{M}(x, z)\ar[d, "\pi"] \\ & \mathcal{L}(x, z) \\ \end{tikzcd} \end{center} is an embedding that satisfies \[ \hat\psi(\rho) \converges{\rho\to\infty}\too (\hat u, \hat v) \in \bar{\mathcal{L}}(x, z) .\] \newpage 2. ("Uniqueness") For any sequence $\ts{\ell_n}\subseteq \mathcal{L}(x, z)$, \[ \ell_n \converges{n\to\infty}\too (\hat u, \hat v) \quad\implies\quad \ell_n \in \im(\hat \psi) \text{ for } n \gg 0 .\] ::: - We already know that $\bar{\mathcal{L}}(x, z)$ is compact and $\mathcal{L}(x, z)$ is a 1-manifold, so we look at neighborhoods of boundary points. - Why unique: will show that the broken trajectory $(\hat u, \hat v)$ is the endpoint of an embedded interval in $\bar{\mathcal{L}}(x, z)$. - Then show that any other sequence converging to $(\hat u, \hat v)$ must approach via this interval, otherwise could have cuspidal points:  $\qed$ \newpage # Gluing Theorem Broken into three steps: 1. **Pre-gluing**: - Get a function $w_\rho$ which interpolates between $u$ and $v$ in the parameter $\rho$. - Not exactly a solution itself, just an "approximation". 2. **Newton's Method**: - Apply the Newton-Picard method to $w_p$ to construct a true solution \[ \psi: [-\rho, \infty) &\to \mathcal{M}(x, z) \\ \rho &\mapsto \oldexp_{w_p}\qty{\gamma(p)} \\ \\ \text{for some } \gamma(p) &\in W^{1, p}(w_p^*\, TW) = T_{w_p} \mathcal{P}(x, z) \] - [GIF of Newton's Method](https://www.maplesoft.com/support/help/content/4702/plot552.gif) 3. **Project and Verify Properties**: - Check that the projection $\hat \psi = \pi \circ \psi$ satisfies the conditions from the theorem. $\qed$ \newpage # 9.3: Pre-gluing, Construction of $w_\rho$ - Choose (once and for all) a bump function $\beta$ on $B_{\eps}(0)^c \subset \RR \to [0, 1]$ which is 1 on $\abs{x} \geq 1$ and $0$ on $\abs{x} < \eps$ - Split into positive and negative parts $\beta^\pm(s)$:  - Define an interpolation $w_\rho$ from $u$ to $v$ in the following way: let - $\oldexp \left[ \wait \right] \da \oldexp_{y(t)}(\wait)$ and - $\ln(\wait) \da \oldexp\inv_{y(t)}(\wait)$, then \[ w_\rho: x &\to z \\ w_\rho(s,t) &\da \begin{cases} {\color{blue} u(s+\rho, t)} & s\in (-\infty, -1] \\ \\ \oldexp\left[ \beta^-(s)\ln({\color{blue} u(s+\rho, t) }) + \beta^+(s)\ln({\color{purple} u(s-\rho, t)} )\right] & s\in [-1, 1] \\ \\ {\color{purple} u(s-\rho, t)} & s\in [1, \infty) \end{cases} .\] - Why does this make sense? \[ \abs{s}\leq 1 \implies u(s\pm \rho, t) \in \ts{\oldexp_{y(t)} Y(t) \st \sup_{t\in S^1} \norm{Y(t)}\leq r_0 } \subseteq \im \oldexp_{y(t)} (\wait) ,\] so we can apply $\oldexp_{y(t)}\inv(\wait)$. - Can make $\abs{s}\leq 1$ for large $\rho$, since \[ u(s, t) \converges{s\to\infty}\too \quad &y(t) \\ v(s, t) \converges{s\to-\infty}\too \quad &y(t) .\] - So pick a $\rho_0$ such that this holds for $\rho > \rho_0$. - Might have to increase $\rho_0$ later in the proof, so $\rho > \rho_0$ just means $\rho \gg 0$. - Some properties: - $w_\rho \in C^\infty(x, z)$ and is differentiable in $\rho$. - $s\in [-\eps, \eps] \implies w_\rho(s, t) = y(t)$. \[ w_\rho(s-\rho, t) &\converges{\rho\to\infty}\too u(s, t)\qtext{in} C^\infty_\loc \\ \\ w_\rho(s, t) &\converges{\rho\to\infty}\too y(t) \qtext{in} C^\infty_\loc .\] - Now carry out the linearized version on tangent vectors, to which we will apply Newton-Picard: - Let $Y\in T_u \mathcal{P}(x, y)$ - Let $Z\in T_v \mathcal{P}(x, y)$ - Replace $w_\rho$ with the interpolation \[ Y\size_\rho Z \in T_{w_\rho} \mathcal{P}(x, y) = W^{1, p}(w_\rho^* TW) .\] defined by \[ (Y\size_\rho Z) (s, t) = \begin{cases} {\color{blue} Y(s+\rho, t)} & s\in (-\infty, -1] \\ \\ \oldexp_T\left[ \beta^-(s) \ln_T({\color{blue} Y(s+\rho, t) }) \, + \beta^+(s) \ln_T({\color{purple} Z(s-\rho, t)} )\right] & s\in [-1, 1] \\ \\ {\color{purple} Z(s-\rho, t)} & s\in [1, \infty) \end{cases} ,\] where the subscript $T$ indicates taking tangents of the exponential maps at appropriate points. $\qed$ \newpage # 9.4: Construction of $\psi$. ## Summary - Newton-Picard method, general idea: - Allows finding zeros of $f$ given an approximate zero $x_0$, using the extra information of the 1st derivative $f'$. - Original method and variant: find the limit of a sequence \[ x_{n+1} = x_n - {f(x_n) \over f'(x_n)} ,\qquad x_{n+1} = x_n - {f(x_n) \over f'({\color{red} x_0} )} .\] - Second variant more useful: only need derivative at one point:  \newpage - Pregluing function $w_\rho \in C^\infty_{\searrow}(x, z)$ from previous section - Exponential decay - Want to construct true solution $\psi_\rho \in \mathcal{M}(x, z)$, so $\mathcal{F}(\psi_p) = 0$. - Suffices to get a weak solution - Automatic continuity + elliptic regularity $\implies$ strong solution - Define $\mathcal{F}_\rho$ as $\mathcal{F} \circ \oldexp_{w_\rho}$ expanded bases $Z_i$ from trivialization of $TW$. - $L_\rho = (d\mathcal{F}_\rho)_0$ will be the linearization of the Floer operator at zero. \vspace{5em} - Adapting Newton-Picard to operators: - $L_\rho$ won't be invertible on entire space, but \[ {1\over f'(x_0)} \iff L_\rho^{-1}, \] - Decompose \[ T_{w_\rho} \mathcal{P}(x, z) = W^{1, p}(w_\rho^* TW) = W^{1, p}(\RR\cross S^1; \RR^{2n}) = \ker(L_{\rho}) \oplus W_{\rho}\perp, \] where $L_\rho$ will have a right inverse on $W_\rho\perp$. - Where does $W_\rho\perp$ come from? Essentially the kernel of some linear functional given by an integral: \[ W_\rho\perp \da \ts{Y\in W^{1, p} \st \int_{\RR\cross S^1} \inner{Y}{\cdots}\,ds\,dt = 0,\, \text{ plus conditions}} .\] - Run Newton-Picard in $W_{\rho}\perp$ - Will obtain for every $\rho \geq \rho_0$ an element $\gamma(\rho) \in W_\rho\perp$ with \[ \mathcal{F}_\rho(\gamma(\rho)) = 0 .\] \newpage - Where does $\gamma$ come from? Intersection-theoretic interpretation on page 320: \[ \qty{\oldexp_{w_\rho}}\inv \mathcal{M}(x, z) \intersect W_{\rho}\perp &\subseteq T_{w_\rho}\mathcal{P}(x, z) &\leadsto \gamma \\ \mathcal{M}(x, z) \cap \ts{\oldexp_{w_\rho} W_\rho\perp \st \rho\geq \rho_0}&\subseteq \mathcal{P}(x, z) &\leadsto \psi(\rho) ,\] which we get by exponentiating. - This gives a codimension 1 subspace in $\mathcal{M}(x, z)$, which we take to be $\psi(\rho)$:  \newpage > Schematic picture here for $\gamma, \psi(\rho)$. \newpage - Apply the implicit function theorem to show differentiability of $\gamma$ in $\rho$. - Use a trivialization $Z_i^\rho$ of $TW$ to get a vector field along $w_\rho$ - This is exactly an element of $T_{w_\rho}\mathcal{P}(x, z)$ - Exponentiate to get an element of $\mathcal{M}(x, z)$: \[ \psi(\rho) \da \oldexp_{w_\rho}\qty{\gamma(\rho)} .\] - **Final Result**: project onto $\mathcal{L}(x, z)$ to get $\hat \psi$. \vspace{4em} **Checking Properties**: - Existence: show $\hat \psi$ is a proper injective immersion $\implies$ embedding. - Uniqueness: show the broken trajectory $(\hat u, \hat v)$ is the endpoint of an embedded interval in $\bar{\mathcal{L}}(x, z)$. - Show that any other sequence converging to $(\hat u, \hat v)$ must approach via this interval, otherwise could have cuspidal points:  > Probably not worth going farther than this! Extremely detailed analysis. $\qed$ \newpage