From cf4dec33e62829f40240f29b484ea2f4c7c401e7 Mon Sep 17 00:00:00 2001 From: Jaron Kent-Dobias Date: Mon, 26 Aug 2024 17:24:53 +0200 Subject: Lots of writing and figure work. --- figs/action_1.pdf | Bin 20655 -> 20905 bytes figs/action_3.pdf | Bin 16245 -> 16245 bytes figs/dynamics_2.pdf | Bin 19291 -> 17774 bytes figs/dynamics_3.pdf | Bin 22214 -> 21696 bytes figs/phases_1.pdf | Bin 11318 -> 11318 bytes figs/phases_2.pdf | Bin 12012 -> 12012 bytes figs/phases_3.pdf | Bin 13050 -> 13049 bytes topology.bib | 14 ++ topology.tex | 605 ++++++++++++++++++++++++++++++---------------------- 9 files changed, 368 insertions(+), 251 deletions(-) diff --git a/figs/action_1.pdf b/figs/action_1.pdf index e4519a4..2a87561 100644 Binary files a/figs/action_1.pdf and b/figs/action_1.pdf differ diff --git a/figs/action_3.pdf b/figs/action_3.pdf index 749088d..54825b1 100644 Binary files a/figs/action_3.pdf and b/figs/action_3.pdf differ diff --git a/figs/dynamics_2.pdf b/figs/dynamics_2.pdf index 41ffece..a0fbcea 100644 Binary files a/figs/dynamics_2.pdf and b/figs/dynamics_2.pdf differ diff --git a/figs/dynamics_3.pdf b/figs/dynamics_3.pdf index b1763e4..b562de4 100644 Binary files a/figs/dynamics_3.pdf and b/figs/dynamics_3.pdf differ diff --git a/figs/phases_1.pdf b/figs/phases_1.pdf index 8f6c482..4a8c9c9 100644 Binary files a/figs/phases_1.pdf and b/figs/phases_1.pdf differ diff --git a/figs/phases_2.pdf b/figs/phases_2.pdf index bf16e76..a73666e 100644 Binary files a/figs/phases_2.pdf and b/figs/phases_2.pdf differ diff --git a/figs/phases_3.pdf b/figs/phases_3.pdf index 506ca73..e7a55fb 100644 Binary files a/figs/phases_3.pdf and b/figs/phases_3.pdf differ diff --git a/topology.bib b/topology.bib index 316df0f..8c17a2c 100644 --- a/topology.bib +++ b/topology.bib @@ -264,6 +264,20 @@ eprinttype = {arxiv} } +@article{Kent-Dobias_2023_How, + author = {Kent-Dobias, Jaron and Kurchan, Jorge}, + title = {How to count in hierarchical landscapes: a full solution to mean-field complexity}, + journal = {Physical Review E}, + publisher = {American Physical Society (APS)}, + year = {2023}, + month = {6}, + number = {6}, + volume = {107}, + pages = {064111}, + url = {https://doi.org/10.1103/PhysRevE.107.064111}, + doi = {10.1103/PhysRevE.107.064111} +} + @article{Kent-Dobias_2023_When, author = {Kent-Dobias, Jaron}, title = {When is the average number of saddle points typical?}, diff --git a/topology.tex b/topology.tex index 102dc6b..a7b68dc 100644 --- a/topology.tex +++ b/topology.tex @@ -34,7 +34,7 @@ \begin{abstract} We consider the set of solutions to $M$ random polynomial equations with - independent Gaussian coefficients on the $(N-1)$-sphere. When solutions + independent Gaussian coefficients and a target value $V_0$ on the $(N-1)$-sphere. When solutions exist, they form a manifold. We compute the average Euler characteristic of this manifold in the limit of large $N$, and find different behavior depending on the ratio $\alpha=M/N$. When $\alpha<\alpha_\text{onset}$, the @@ -110,7 +110,7 @@ tissues \cite{Fyodorov_2019_A, Fyodorov_2020_Counting, Montanari_2024_On, Kent-Dobias_2024_Conditioning, Kent-Dobias_2024_Algorithm-independent}. In each of these cases, the authors studied properties of the cost function \begin{equation} \label{eq:cost} - \mathscr C(\mathbf x)=\frac12\sum_{k=1}^MV_k(\mathbf x)^2 + \mathscr C(\mathbf x)=\frac12\sum_{k=1}^M\big[V_k(\mathbf x)-V_0\big]^2 \end{equation} which achieves zero only for configurations that satisfy all the constraints. Here we dispense with the cost function and study the set of solutions @@ -136,9 +136,188 @@ direct sum of the Betti numbers of $\Omega$. We find that for the varied cases we study, these two always coincide at the largest exponential order in $N$, putting strong constraints on the resulting topology and geometry. -\section{Results} +\section{Methods} -\subsection{Topology of solutions to many equations and the satisfiability transition} +\subsection{The average Euler characteristic} + +The Euler characteristic $\chi$ of a manifold is a topological invariant \cite{Hatcher_2002_Algebraic}. It is +perhaps most familiar in the context of connected compact orientable surfaces, where it +characterizes the number of handles in the surface: $\chi=2(1-\#)$ for $\#$ +handles. For general $d$, the Euler characteristic of the $d$-sphere is $2$ if $d$ is even and 0 if $d$ is odd. The canonical method for computing the Euler characteristic is done by +defining a complex on the manifold in question, essentially a +higher-dimensional generalization of a polygonal tiling. Then $\chi$ is given +by an alternating sum over the number of cells of increasing dimension, which +for 2-manifolds corresponds to the number of vertices, minus the number of +edges, plus the number of faces. + +Morse theory offers another way to compute the Euler characteristic using the +statistics of stationary points of a function $H:\Omega\to\mathbb R$ \cite{Audin_2014_Morse}. For +functions $H$ without any symmetries with respect to the manifold, the surfaces +of gradient flow between adjacent stationary points form a complex. The +alternating sum over cells to compute $\chi$ becomes an alternating sum over +the count of stationary points of $H$ with increasing index, or +\begin{equation} + \chi=\sum_{i=0}^N(-1)^i\mathcal N_H(\text{index}=i) +\end{equation} +Conveniently, we can express this abstract sum as an integral over the manifold +using a small variation on the Kac--Rice formula for counting stationary +points. Since the sign of the determinant of the Hessian matrix of $H$ at a +stationary point is equal to its index, if we count stationary points including +the sign of the determinant, we arrive at the Euler characteristic, or +\begin{equation} \label{eq:kac-rice} + \chi(\Omega)=\int_\Omega d\mathbf x\,\delta\big(\nabla H(\mathbf x)\big)\det\operatorname{Hess}H(\mathbf x) +\end{equation} +When the Kac--Rice formula is used to \emph{count} stationary points, the sign +of the determinant is a nuisance that one must take pains to preserve +\cite{Fyodorov_2004_Complexity}. Here we are correct to exclude it. + +We need to choose a function $H$ for our calculation. Because $\chi$ is +a topological invariant, any choice will work so long as it does not share some +symmetry with the underlying manifold, i.e., that it $H$ satisfies the Smale condition. Because our manifold of random +constraints has no symmetries, we can take a simple height function $H(\mathbf +x)=\mathbf x_0\cdot\mathbf x$ for some $\mathbf x_0\in\mathbb R^N$ with +$\|\mathbf x_0\|^2=N$. $H$ is a height function because when $\mathbf x_0$ is +used as the polar axis, $H$ gives the height on the sphere. + +We treat the integral over the implicitly defined manifold $\Omega$ using the +method of Lagrange multipliers. We introduce one multiplier $\omega_0$ to +enforce the spherical constraint and $M$ multipliers $\omega_k$ to enforce the vanishing of +each of the $V_k$, resulting in the Lagrangian +\begin{equation} \label{eq:lagrangian} + L(\mathbf x,\pmb\omega) + =H(\mathbf x)+\frac12\omega_0\big(\|\mathbf x\|^2-N\big) + +\sum_{k=1}^M\omega_k\big(V_k(\mathbf x)-V_0\big) +\end{equation} +The integral over $\Omega$ in \eqref{eq:kac-rice} then becomes +\begin{equation} \label{eq:kac-rice.lagrange} + \chi(\Omega)=\int_{\mathbb R^N} d\mathbf x\int_{\mathbb R^{M+1}}d\pmb\omega + \,\delta\big(\partial L(\mathbf x,\pmb\omega)\big) + \det\partial\partial L(\mathbf x,\pmb\omega) +\end{equation} +where $\partial=[\frac\partial{\partial\mathbf x},\frac\partial{\partial\pmb\omega}]$ +is the vector of partial derivatives with respect to all $N+M+1$ variables. +This integral is now in a form where standard techniques from mean-field theory +can be applied to calculate it. + +In order for certain Gaussian integrals in the following calculation to be +well-defined, it is necessary to treat instead the Lagrangian problem above +with $\pmb\omega\mapsto i\pmb\omega$. This transformation does not effect the +Dirac $\delta$ functions of the gradient, but it does change the determinant by +a factor of $i^{N+M+1}$. We will see that the result of the rest of the +calculation neglecting this factor is real. Since the Euler characteristic is +also necessarily real, this indicates an inconsistency with this transformation +when $N+M+1$ is odd. In fact, the Euler characteristic is always zero for +odd-dimensional manifolds. This is the signature of it in this problem. + +\subsubsection{Calculation of the average Euler characteristic} + +To evaluate the average of $\chi$ over the constraints, we first translate the $\delta$ functions and determinant to integral form, with +\begin{align} + \delta\big(\partial L(\mathbf x,\pmb\omega)\big) + &=\int\frac{d\hat{\mathbf x}}{(2\pi)^N}\frac{d\hat{\pmb\omega}}{(2\pi)^{M+1}} + e^{i[\hat{\mathbf x},\hat{\pmb\omega}]\cdot\partial L(\mathbf x,\pmb\omega)} + \\ + \det\partial\partial L(\mathbf x,\pmb\omega) + &=\int d\bar{\pmb\eta}\,d\pmb\eta\,d\bar{\pmb\gamma}\,d\pmb\gamma\, + e^{-[\bar{\pmb\eta},\bar{\pmb\gamma}]^T\partial\partial L(\mathbf x,\pmb\omega)[\pmb\eta,\pmb\gamma]} +\end{align} +where $\hat{\mathbf x}$ and $\hat{\pmb\omega}$ are ordinary vectors and +$\bar{\pmb\eta}$, $\pmb\eta$, $\bar{\pmb\gamma}$, and $\pmb\gamma$ are +Grassmann vectors. With these expressions substituted into +\eqref{eq:kac-rice.lagrange}, the result is a integral over an exponential +whose argument is linear in the random functions $V_k$. These functions can +therefore be averaged over, and the resulting expression treated with standard +methods. Details of this calculation can be found in Appendix~\ref{sec:euler}. +The result is the reduction of the average Euler characteristic to an expression of the +form +\begin{equation} + \overline{\chi(\Omega)} + =\int dR\,dD\,dm\,d\hat m\,g(R,D,m,\hat m)\,e^{N\mathcal S_\Omega(R,D,m,\hat m)} +\end{equation} +where $g$ is a prefactor subexponential in $N$, and $\mathcal S_\Omega$ is an effective action defined by +\begin{equation} + \begin{aligned} + \mathcal S_\Omega(R,D,m,\hat m\mid\alpha,V_0) + &=\hat m-\frac\alpha2\left[ + \log\left(1+\frac{f(1)D}{f'(1)R^2}\right) + +\frac{V_0^2}{f(1)}\left(1+\frac{f'(1)R^2}{f(1)D}\right)^{-1} + \right] \\ + &\hspace{7em}+\frac12\log\left( + 1+\frac{(1-m^2)D+\hat m^2-2Rm\hat m}{R^2} + \right) + \end{aligned} +\end{equation} +The remaining order parameters defined by the scalar products +\begin{align} + R=-i\frac1N\mathbf x\cdot\hat{\mathbf x} + && + D=\frac1N\hat{\mathbf x}\cdot\hat{\mathbf x} + && + m=\frac1N\mathbf x\cdot\mathbf x_0 + && + \hat m=-i\frac1N\hat {\mathbf x}\cdot\mathbf x_0 +\end{align} + +This integral can be evaluated by a saddle point method. For reasons we will +see, it is best to extremize with respect to $R$, $D$, and $\hat m$, leaving a +new effective action of $m$ alone. This can be solved to give +\begin{equation} + D=-\frac{m+R^*(m)}{1-m^2} \qquad \hat m=0 +\end{equation} +\begin{equation} + \begin{aligned} + R^*(m) + &=\frac{-m(1-m^2)}{2[f(1)-(1-m^2)f'(1)]^2} + \Bigg[ + \alpha V_0^2f'(1) + +(2-\alpha)f(1)\left(\frac{f(1)}{1-m^2}-f'(1)\right) \\ + &\quad+\operatorname{sgn}(m)\alpha\sqrt{ + \tfrac{4V_0^2}\alpha f(1)f'(1)\left[\tfrac{f(1)}{1-m^2}-f'(1)\right] + +\left[\tfrac{f(1)^2}{1-m^2}-\big(V_0^2+f(1)\big)f'(1)\right]^2 + } + \Bigg] + \end{aligned} +\end{equation} +\begin{equation} + \mathcal S_\Omega(m) + =-\frac\alpha2\bigg[ + \log\left( + 1-\frac{f(1)}{f'(1)}\frac{1+m/R^*}{1-m^2} + \right) + +\frac{V_0^2}{f(1)}\left( + 1-\frac{f'(1)}{f(1)}\frac{1-m^2}{1+m/R^*} + \right)^{-1} + \bigg] + +\frac12\log\left(-\frac m{R^*}\right) +\end{equation} +To finish evaluating the integral, this expression should be maximized with +respect to $m$. The order parameter $m$ is both physical and interpretable, as +it gives the overlap of the configuration $\mathbf x$ with the height axis +$\mathbf x_0$. Therefore, the value $m^*$ that maximizes this action can be +understood as the latitude on the sphere where most of the contribution to the +Euler characteristic is made. + +The action $\mathcal S_\Omega$ is extremized with respect to $m$ at $m^*=0$ or $m^*=-R^*$. +In the latter case, $m^*$ takes the value +\begin{equation} + m^*=\pm\sqrt{1-\frac{\alpha}{f'(1)}\big(V_0^2+f(1)\big)} +\end{equation} +and $\mathcal S_\Omega(m^*)=0$. However, when +\begin{equation} + V_0^2>V_\text{on}^2\equiv\frac{1-\alpha+\sqrt{1-\alpha}}\alpha f(1) +\end{equation} +$R^*(m^*)$ becomes complex and the solution is no longer valid. Likewise, when +\begin{equation} + V_0^2