\documentclass{SciPost} % Prevent all line breaks in inline equations. \binoppenalty=10000 \relpenalty=10000 \usepackage[utf8]{inputenc} \usepackage[T1]{fontenc} \usepackage{amsmath,latexsym,graphicx} \usepackage[bitstream-charter]{mathdesign} \urlstyle{sf} \fancypagestyle{SPstyle}{ \fancyhf{} \lhead{\colorbox{scipostblue}{\bf \color{white} ~SciPost Physics }} \rhead{{\bf \color{scipostdeepblue} ~Submission }} \renewcommand{\headrulewidth}{1pt} \fancyfoot[C]{\textbf{\thepage}} } % Fix \cal and \mathcal characters look (so it's not the same as \mathscr) \DeclareSymbolFont{usualmathcal}{OMS}{cmsy}{m}{n} \DeclareSymbolFontAlphabet{\mathcal}{usualmathcal} \hypersetup{ colorlinks, linkcolor={red!50!black}, citecolor={blue!50!black}, urlcolor={blue!80!black} } \begin{document} \pagestyle{SPstyle} \begin{center}{\Large \textbf{\color{scipostdeepblue}{ On the topology of solutions to random continuous constraint satisfaction problems\\ }}}\end{center} \begin{center} \textbf{Jaron Kent-Dobias} \end{center} \begin{center} Istituto Nazionale di Fisica Nucleare, Sezione di Roma I, Italy \\ ICTP South American Institute for Fundamental Research, São Paulo, Brazil \\ Instituto de Física Teórica, Universidade Estadual Paulista ``Júlio de Mesquita Filho'', São Paulo, Brazil \\[\baselineskip] \href{mailto:jaron@ictp-saifr.org}{\small jaron@ictp-saifr.org} \end{center} \section*{\color{scipostdeepblue}{Abstract}} \textbf{\boldmath{% We consider the set of solutions to $M$ random polynomial equations whose $N$ variables are restricted to the $(N-1)$-sphere. Each equation has independent Gaussian coefficients and a target value $V_0$. 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 target value $V_0$, the ratio $\alpha=M/N$, and the variances of the coefficients. We divide this behavior into five phases with different implications for the topology of the solution manifold. When $M=1$ there is a correspondence between this problem and level sets of the energy in the spherical spin glasses. We conjecture that the transition energy dividing two of the topological phases corresponds to the energy asymptotically reached by gradient descent from a random initial condition, possibly resolving an open problem in out-of-equilibrium dynamics. However, the quality of the available data leaves the question open for now. } } \vspace{\baselineskip} %%%%%%%%%% BLOCK: Copyright information % This block will be filled during the proof stage, and finilized just before publication. % It exists here only as a placeholder, and should not be modified by authors. \noindent\textcolor{white!90!black}{% \fbox{\parbox{0.975\linewidth}{% \textcolor{white!40!black}{\begin{tabular}{lr}% \begin{minipage}{0.6\textwidth}% {\small Copyright attribution to authors. \newline This work is a submission to SciPost Physics. \newline License information to appear upon publication. \newline Publication information to appear upon publication.} \end{minipage} & \begin{minipage}{0.4\textwidth} {\small Received Date \newline Accepted Date \newline Published Date}% \end{minipage} \end{tabular}} }} } %%%%%%%%%% BLOCK: Copyright information %%%%%%%%%% TODO: LINENO % For convenience during refereeing we turn on line numbers: %\linenumbers % You should run LaTeX twice in order for the line numbers to appear. %%%%%%%%%% END TODO: LINENO %%%%%%%%%% TODO: TOC % Guideline: if your paper is longer that 6 pages, include a TOC % To remove the TOC, simply cut the following block \vspace{10pt} \noindent\rule{\textwidth}{1pt} \tableofcontents \noindent\rule{\textwidth}{1pt} \vspace{10pt} %%%%%%%%%% END TODO: TOC \section{Introduction} Constraint satisfaction problems seek configurations that simultaneously satisfy a set of equations, and form a basis for thinking about problems as diverse as neural networks \cite{Mezard_2009_Constraint}, granular materials \cite{Franz_2017_Universality}, ecosystems \cite{Altieri_2019_Constraint}, and confluent tissues \cite{Urbani_2023_A}. All but the last of these examples deal with sets of inequalities, while the last considers a set of equality constraints. Inequality constraints are familiar in situations like zero-cost solutions in neural networks with ReLu activations and stable equilibrium in the forces between physical objects. Equality constraints naturally appear in the zero-gradient solutions to overparameterized smooth neural networks and in vertex models of tissues. In problems ranging from toy models \cite{Baldassi_2016_Unreasonable, Baldassi_2019_Properties} to real deep neural networks \cite{Goodfellow_2014_Qualitatively, Draxler_2018_Essentially, Frankle_2020_Revisiting, Vlaar_2022_What, Wang_2023_Plateau}, there is great interest in characterizing structure in the set of solutions, which can influence the behavior of algorithms trying to find them \cite{Beneventano_2023_On}. Here, we show how topological information about the set of solutions can be calculated in a simple problem of satisfying random nonlinear equalities. This allows us to reason about the connectivity and structure of the solution set. The topological properties revealed by this calculation yield surprising results for the well-studied spherical spin glasses, where a topological transition thought to occur at a threshold energy $E_\text{th}$ where marginal minima are dominant is shown to occur at a different energy $E_\text{sh}$. We conjecture that this difference resolves an outstanding problem with the out-of-equilibrium dynamics in these systems. We consider the problem of finding configurations $\mathbf x\in\mathbb R^N$ lying on the $(N-1)$-sphere $\|\mathbf x\|^2=N$ that simultaneously satisfy $M$ nonlinear constraints $V_k(\mathbf x)=V_0$ for $1\leq k\leq M$ and some constant $V_0\in\mathbb R$. The nonlinear constraints are taken to be centered Gaussian random functions with covariance \begin{equation} \label{eq:covariance} \overline{V_i(\mathbf x)V_j(\mathbf x')} =\delta_{ij}f\left(\frac{\mathbf x\cdot\mathbf x'}N\right) \end{equation} for some choice of function $f$. When the covariance function $f$ is polynomial, the $V_k$ are also polynomial, with a term of degree $p$ in $f$ corresponding to all possible terms of degree $p$ in the $V_k$. One can explicitly construct functions that satisfy \eqref{eq:covariance} by taking \begin{equation} V_k(\mathbf x) =\sum_{p=0}^\infty\frac1{p!}\sqrt{\frac{f^{(p)}(0)}{N^p}} \sum_{i_1\cdots i_p}^NJ^{(k,p)}_{i_1\cdots i_p}x_{i_1}\cdots x_{i_p} \end{equation} where the elements of the tensors $J^{(k,p)}$ are independently distributed unit normal random variables. The series coefficients of $f$ therefore control the variances of the random coefficients in the polynomials $V_k$. When $M=1$, this problem corresponds to finding the level set of a spherical spin glass at energy density $E=V_0/\sqrt{N}$. This problem or small variations thereof have attracted attention recently for their resemblance to encryption, least-squares optimization, and vertex models of confluent tissues \cite{Fyodorov_2019_A, Fyodorov_2020_Counting, Fyodorov_2022_Optimization, Tublin_2022_A, Vivo_2024_Random, Urbani_2023_A, Kamali_2023_Dynamical, Kamali_2023_Stochastic, Urbani_2024_Statistical, Montanari_2023_Solving, 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}^M\big[V_k(\mathbf x)-V_0\big]^2 \end{equation} which achieves zero only for configurations that satisfy all the constraints. Introduced in Ref.~\cite{Fyodorov_2019_A}, the existence of solutions and the geometric structure of the cost function were studied for the problem with linear $V_k$ in a series of papers \cite{Fyodorov_2019_A, Fyodorov_2020_Counting, Fyodorov_2022_Optimization} and later reviewed \cite{Vivo_2024_Random}. Some work on the equilibrium measure of the cost function with nonlinear $V_k$ was made in Ref.~\cite{Tublin_2022_A}, and the problem was solved in Ref.~\cite{Urbani_2023_A}. Subsequent work has studied varied dynamics applied to the cost function, including gradient descent, Hessian descent, Langevin, stochastic gradient descent, and approximate message passing \cite{Kamali_2023_Dynamical, Kamali_2023_Stochastic, Montanari_2023_Solving, Montanari_2024_On}. Finally, some progress has been made on aspects of the geometric structure of the cost function with nonlinear $V_k$ \cite{Kent-Dobias_2024_Conditioning, Kent-Dobias_2024_Algorithm-independent}. From the perspective of the cost function, the set of solutions looks like a network of flat canyons at the bottom of the cost landscape. Here we dispense with the cost function and study the set of solutions directly. This set can be written as \begin{equation} \Omega=\big\{\mathbf x\in\mathbb R^N\mid \|\mathbf x\|^2=N,V_k(\mathbf x)=V_0 \;\forall\;k=1,\ldots,M\big\} \end{equation} Because the constraints are all smooth functions, $\Omega$ is almost always a manifold without singular points.\footnote{The conditions for a singular point are that $0=\frac\partial{\partial\mathbf x}V_k(\mathbf x)$ for all $k$. This is equivalent to asking that the constraints $V_k$ all have a stationary point at the same place. When the $V_k$ are independent and random, this is vanishingly unlikely, requiring $NM+1$ independent equations to be simultaneously satisfied. This means that different connected components of the set of solutions do not intersect, nor are there self-intersections, without extraordinary fine-tuning.} We study the topology of the manifold $\Omega$ by computing its average Euler characteristic, a topological invariant whose value puts constraints on the manifold's structure. The topological phases determined by this measurement are distinguished by the size and sign of the Euler characteristic, and the distribution in space of its constituents. In Section~\ref{sec:euler.1} we describe how to calculate the average Euler characteristic, how to interpret the results of that calculation, and what topological phases are implied. In Section~\ref{sec:ssg} we examine some implications of these results for dynamic thresholds in the spherical spin glasses. Finally, in Section~\ref{sec:conclusion} we make some concluding remarks. Many of the details of the calculations in the middle sections are found in Appendices~A--D. \section{The average Euler characteristic} \label{sec:euler.1} \subsection{Definition and derivation} 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. In higher dimensions it is more difficult to interpret, but there are a few basic intuitions. The Euler characteristic of the hypersphere is $2$ in even dimensions and 0 in odd dimensions. In fact, the Euler characteristic of an odd-dimensional manifold is always zero. The Euler characteristic of the union of two disjoint manifolds is the sum of the Euler characteristics of the individual manifolds, and that of the product of two manifolds is the product of the Euler characteristics. This means that a manifold made of many disconnected sphere-like components will have a large positive Euler characteristic. A manifold with many hyper-handles will have a large negative Euler characteristic. And no matter the Euler characteristic of a manifold, the Euler characteristic of its product with the circle $S^1$ is zero. The canonical method for computing the Euler characteristic is to construct a complex on the manifold in question, which is 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 of a manifold $\Omega$ using the statistics of stationary points in 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 becomes an alternating sum over the count of stationary points of $H$ with increasing index, or \begin{equation} \chi(\Omega)=\sum_{i=0}^N(-1)^i\mathcal N_H(\text{index}=i) \end{equation} Conveniently, we can express this sum as an integral over the manifold using a small variation on the Kac--Rice formula for counting stationary points \cite{Kac_1943_On, Rice_1939_The}. 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 calculate the total number stationary points, one must take pains to eliminate the sign of the determinant \cite{Fyodorov_2004_Complexity}. Here it is correct to preserve 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 have degenerate stationary points on the manifold, i.e., that it is a Morse function, and that it does not share some symmetry with the underlying manifold, i.e., that it satisfies the Smale condition. Because our manifold is random and 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 $\|\mathbf x_0\|^2=N$. We call $H$ a height function because when $\mathbf x_0$ is interpreted as the polar axis of a spherical coordinate system, $H$ gives the height on the sphere relative to the equator. 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 $\omega_k$ for $k=1,\ldots,M$ to enforce the $M$ constraints, 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 the solution manifold $\Omega$ in \eqref{eq:kac-rice} 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 expression is now in a form where standard techniques from the mean-field theory of disordered systems can be applied to average over the random constraint functions and evaluate the integrals to leading order in large $N$. Details of this calculation can be found in Appendix~\ref{sec:euler}. The result is the reduction of the average Euler characteristic to an integral over a single order parameter $m=\frac1N\mathbf x\cdot\mathbf x_0$ of the form \begin{equation} \overline{\chi(\Omega)}=\left(\frac{N}{2\pi}\right)^{\frac12}\int dm\,g(m)\,e^{N\mathcal S_\chi(m)} \end{equation} where $g(m)$ is a prefactor of order $N^0$ and $\mathcal S_\chi(m)$ is an effective action defined by \begin{equation} \label{eq:S.m} \mathcal S_\chi(m) =-\frac\alpha2\bigg[ \log\left( 1-\frac{f(1)}{f'(1)}\frac{1+\frac m{R_m}}{1-m^2} \right) +\frac{V_0^2}{f(1)}\left( 1-\frac{f'(1)}{f(1)}\frac{1-m^2}{1+\frac m{R_m}} \right)^{-1} \bigg] +\frac12\log\left(-\frac m{R_m}\right) \end{equation} Here we have introduced the ratio $\alpha=M/N$ between the number of equations and the number of variables, and $R_m$ is a function of $m$ given by \begin{equation} \label{eq:rs} \begin{aligned} R_m \equiv\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 \\ \quad+\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} The effective action \eqref{eq:S.m} is plotted in Fig.~\ref{fig:action} for a selection of parameters. To finish evaluating the integral by the saddle-point approximation, the action should be maximized with respect to $m$. If $m_*$ is such a maximum, then the resulting average Euler characteristic is $\overline{\chi(\Omega)}\propto e^{N\mathcal S_\chi(m_*)}$. In the next subsection we examine the maxima of $\mathcal S_\chi$ and their properties as the parameters are varied. \begin{figure}[tb] \includegraphics{figs/action_1.pdf} \hspace{-3.5em} \includegraphics{figs/action_3.pdf} \caption{ \textbf{Effective action for the Euler characteristic.} The action \eqref{eq:S.m} as a function of $m=\frac1N\mathbf x\cdot\mathbf x_0$ for pure polynomial constraints and a selection of target values $V_0$. Dashed lines depict $\operatorname{Re}\mathcal S_\chi$ when its imaginary part is nonzero. In both plots $\alpha=\frac12$. \textbf{Left:} With linear functions there are two regimes. For small $V_0$, there are maxima at $m=\pm m_*$ where the action is zero, while for $V_0> V_{\text{\textsc{sat}}\ast}=1$ the action is negative everywhere. \textbf{Right:} With nonlinear functions there are other possible regimes. For small $V_0$, there are maxima at $m=\pm m_*$ but the real part of the action is maximized at $m=0$ where the action is complex. For larger $V_0\geq V_\text{on}\simeq1.099$ the maxima at $m=\pm m_*$ disappear. For $V_0\geq V_\text{sh}\simeq1.394$ larger still, the action becomes real everywhere. Finally, for $V_0>V_\text{\textsc{sat}}\simeq1.440$ the action is negative everywhere. } \label{fig:action} \end{figure} \subsection{Features of the effective action} The order parameter $m$ is 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 at which most of the contribution to the Euler characteristic is made.\footnote{ The order parameter $m$ may resemble the magnetization that appears in problems that have a signal or spike, where it gives the overlap of a configuration with the hidden signal. Here $\textbf x_0$ is no signal, but a direction chosen uniformly at random and with no significance to the set of solutions. Here, if a feature of the action is present at some value $m$, it should be interpreted as indicating that, with overwhelming probability, typical configurations contributing to that feature have an overlap $m$ with a typical point in configuration space. For instance, for $m$ sufficiently close to 1, $\mathcal S_\chi(m)$ is always negative, which is a result of the absence of any stationary points contributing to the Euler characteristic at those overlaps. Given a random height axis $\mathbf x_0$, the nearest point to $\mathbf x_0$ on the solution manifold will be the absolute maximum of the height function, and therefore will contribute to the Euler characteristic. Hence the region of negative action in the vicinity of $m=1$ implies there is a typical minimum distance between the solution manifold and a randomly drawn point in configuration space, and that it is vanishingly unlikely to draw a point in configuration space uniformly at random and find it any closer to the solution manifold than this. Other properties of the set of solutions could be studied by drawing $\mathbf x_0$ from an alternative distribution, like the Boltzmann distribution of the cost function, from the set of its stationary points, or from the solution manifold itself. While the value of the Euler characteristic would not change, the dependence of the effective action on $m$ would change. } The action $\mathcal S_\chi$ is extremized with respect to $m$ at $m=0$ or at $m=\pm m_*$ for \begin{equation} m_*=\sqrt{1-\frac{\alpha}{f'(1)}\big(V_0^2+f(1)\big)} \end{equation} At these latter extrema, $\mathcal S_\chi(\pm m_*)=0$. Zero action implies that $\overline{\chi(\Omega)}$ does not vary exponentially with $N$, and in fact we show in Appendix~\ref{sec:prefactor} that the contribution from these extrema is $1+o(N^0)$ at $-m_*$ and $(-1)^{N-M-1}+o(N^0)$ at $+m_*$, so that their sum is $2$ in even dimensions and $0$ in odd dimensions. When these extrema exist and maximize the action, this result is consistent with the topology of an $N-M-1$ sphere. If this solution were always well-defined, it would vanish when the argument of the square root vanishes for \begin{equation} V_0^2>V_{\text{\textsc{sat}}\ast}^2\equiv\frac{f'(1)}\alpha-f(1) \end{equation} This corresponds precisely to the satisfiability transition found in previous work by a replica symmetric analysis of the cost function \eqref{eq:cost} \cite{Fyodorov_2019_A, Fyodorov_2020_Counting, Fyodorov_2022_Optimization, Tublin_2022_A, Vivo_2024_Random}. However, the action is not clearly defined in the entire range $m^2<1$: it becomes complex in the region $m^2V_{\text{\textsc{sat}}\ast}^2$, so the naïve satisfiability transition happens first. On the other hand, when $f(q)$ contains powers of $q$ strictly greater than 1, then $f'(1)\geq 2f(1)$ and $V_\text{on}^2\leq V_{\text{\textsc{sat}}\ast}^2$, so the onset happens first. In situations with mixed constant, linear, and nonlinear terms in $f$, the order of the transitions depends on the precise form of $f$. Now we return to the extremum at $m=0$. As for those at $\pm m_*$, the action evaluated at this solution is sometimes complex-valued and sometimes real-valued. For $V_0$ less than a shattering value $V_\text{sh}$ defined by \begin{equation} V_\text{sh}^2\equiv\frac{f(1)}\alpha\left(1-\frac{f(1)}{f'(1)}\right)\left(1+\sqrt{1-\alpha}\right)^2 \end{equation} the maximum at $m=0$ is complex while for $V_0$ greater than this value the action is real. For purely linear $f(q)$, $V_\text{sh}=0$ and the action at $m=0$ is always real, though for $V_0^20$, and $\overline{\chi(\Omega)^2}\neq[\overline{\chi(\Omega)}]^2$ always. } we identify three saddle points that could contribute to the value of $\overline{\chi(\Omega)^2}$: two at $\pm m_*$ where $\frac1N\log\overline{\chi(\Omega)^2}=\frac1N\log\overline{\chi(\Omega)}\simeq0$, and one at $m=0$ where \begin{equation} \frac1N\log\overline{\chi(\Omega)^2}=2\operatorname{Re}\mathcal S_\chi(0) \end{equation} which is consistent with $\overline{\chi(\Omega)^2}\simeq[\overline{\chi(\Omega)}]^2$. We therefore conclude that when the effective action is complex-valued, the average Euler characteristic is negative and its magnitude is given by the real part of the action. Such a correspondence, which indicates that the `annealed' calculation presented here is also representative of typical realizations of the constraints, is not always true. Sometimes the average squared Euler characteristic has alternative saddle points for which $\overline{\chi(\Omega)^2}\neq[\overline{\chi(\Omega)}]^2$, which implies that average properties will not be typical of most realizations. With our calculation of the average squared Euler characteristic, we can identify instabilities of the solution described above toward such replica symmetry breaking (\textsc{rsb}) solutions. The analysis of these instabilities can be found in Appendix~\ref{sec:rsb.instability}. We do not explore these \textsc{rsb} solutions here, except in the context of $M=1$ and the spherical spin glasses in Section~\ref{sec:ssg}. However, in the phase diagrams of Figures \ref{fig:phases} and \ref{fig:crossover} we shade the region where our calculation indicates that an instability is present. \subsection{Topological phases and their interpretation} The results of the previous section allow us to unambiguously define distinct topological phases, which differ depending on the presence or absence of the local maxima at $m=\pm m_*$, on the presence or absence of the local maximum at $m=0$, on the real or complex nature of this maximum, and finally on whether the action is positive or negative. Below we enumerate these regimes, which are schematically represented in Fig.~\ref{fig:cartoons}.\footnote{ In the following we characterize regimes by values of $\overline{\chi(\Omega)}$. These should be understood as their values in \emph{even} dimensions, since in odd dimensions the Euler characteristic is always identically zero. We do not expect the qualitative results to change depending on the evenness or oddness of the manifold dimension. } It is not possible to definitively ascertain what structural features of the solution manifold lead to these average invariants, but we suggest a simplest interpretation consistent with the calculated properties. \begin{figure} \includegraphics[width=0.196\textwidth]{figs/connected.pdf} \includegraphics[width=0.196\textwidth]{figs/middle.pdf} \includegraphics[width=0.196\textwidth]{figs/complex.pdf} \includegraphics[width=0.196\textwidth]{figs/shattered.pdf} \includegraphics[width=0.196\textwidth]{figs/gone.pdf} \hspace{1.5em} \textbf{Regime I} \hfill \textbf{Regime II} \hfill \textbf{Regime III} \hfill \textbf{Regime IV} \hfill \textbf{Regime V} \hspace{1.5em} \caption{ \textbf{Cartoons of the solution manifold in five topological regimes.} The solution manifold is shown as a shaded region, and the height axis $\mathbf x_0$ is a black arrow. In Regime I, the average Euler characteristic is consistent with a manifold with a single simply-connected component. In Regime II, holes occupy the equator but the temperate regions are topologically simple. In Regime III, holes dominate and the edge of the manifold is not necessarily simple. In Regime IV, disconnected components dominate. In Regime V, the manifold is empty. } \label{fig:cartoons} \end{figure} \paragraph{Regime I: \boldmath{$\overline{\chi(\Omega)}=2$}.} This regime is found when the magnitude of the target value $V_0$ is less than the onset $V_\text{on}$ and $\operatorname{Re}\mathcal S(0)<0$, so that the maxima at $m=\pm m_*$ exist and are the dominant contributions to the average Euler characteristic. Here, $\overline{\chi(\Omega)}=2+o(1)$ for even $N-M-1$, strongly indicating a topology homeomorphic to the $S^{N-M-1}$ sphere. This regime is the only nontrivial one found with linear covariance $f(q)=q$, where the solution manifold must be a sphere if it is not empty. \paragraph{Regime II: \boldmath{$\overline{\chi(\Omega)}$} large and negative, isolated contributions at \boldmath{$m=\pm m_*$}.} This regime is found when the magnitude of the target value $V_0$ is less than the onset $V_\text{on}$, $\operatorname{Re}\mathcal S(0)>0$, and the value of the action at $m=0$ is complex. The dominant contribution to the average Euler characteristic comes from the equator at $m=0$, but the complexity of the action implies that the Euler characteristic is negative. While the topology of the manifold is not necessarily connected in this regime, holes are more numerous than components. Since $V_0^2V_\text{on}^2$. The solutions at $m=\pm m_*$ no longer exist, and nontrivial contributions to the Euler characteristic are made all the way to the edges of the solution manifold. \paragraph{Regime IV: \boldmath{$\overline{\chi(\Omega)}$} large and positive.} This regime is found when the magnitude of the target value $V_0$ is greater than the shattering value $V_\text{sh}$ and $\mathcal S(0)>0$. Above the shattering transition the effective action is real everywhere, and its value at the equator is the dominant contribution. Large connected components of the manifold may or may not exist, but small disconnected components outnumber holes.\footnote{ We interpret the large Euler characteristic to indicate a manifold with many (topologically) spherical disconnected components because the manifold is formed by the process of repeatedly taking non-self-intersecting slices of the previous manifold, starting with a sphere. Therefore, an outcome consisting mostly of (topological) spheres seems most plausible. However, a large Euler characteristic is also consistent with a variety of connected product manifolds, among other exotic possibilities. Definitely ruling out such scenarios is not within the scope of this paper. } \paragraph{Regime V: \boldmath{$\overline{\chi(\Omega)}$} very small.} Here $\frac1N\log\overline{\chi(\Omega)}<0$, indicating that the average Euler characteristic shrinks exponentially with $N$. Under most conditions we conclude this is the \textsc{unsat} regime where no manifold exists, but there may be circumstances where part of this regime is characterized by nonempty solution manifolds that are overwhelmingly likely to have Euler characteristic zero. \begin{figure} \includegraphics{figs/phases_1.pdf} \hspace{-3em} \includegraphics{figs/phases_2.pdf} \hspace{-3em} \includegraphics{figs/phases_3.pdf} \caption{ \textbf{Topological phase diagram.} Topological phases of the problem for three different homogeneous covariance functions. The regimes are defined in the text and depicted as cartoons in Fig.~\ref{fig:cartoons}. The shaded region in the center panel shows where these results are unstable to \textsc{rsb}. In the limit of $\alpha\to0$, the behavior of level sets of the spherical spin glasses are recovered: the righthand plot shows how the ground state energy $E_\text{gs}$ and threshold energy $E_\text{th}$ of the 3-spin spherical model correspond with the limits of the satisfiability and shattering transitions in the pure cubic problem. Note that for mixed models with inhomogeneous covariance functions, $E_\text{th}$ is not the lower limit of $V_\text{sh}$. } \label{fig:phases} \end{figure} \paragraph{} The distribution of these phases for situations with homogeneous polynomial constraint functions is shown in Fig.~\ref{fig:phases}. For purely linear models, the only two regimes are I and V, separated by a satisfiability transition at $V_{\text{\textsc{sat}}\ast}$. This is expected: the intersection of a plane and a sphere is another sphere, and therefore a model of linear constraints in a spherical configuration space can only produce a solution manifold consisting of a single sphere, or the empty set. For purely nonlinear models, regime I does not appear, while the other three nontrivial regimes do. Regimes II and III are separated by the onset transition at $V_\text{on}$, while III and IV are separated by the shattering transition at $V_\text{sh}$. Finally, IV and V are now separated by the satisfiability transition at $V_\text{\textsc{sat}}$. An interesting feature occurs in the limit of $\alpha$ to zero. If $V_0$ is likewise rescaled in the correct way, the limit of these phase boundaries approaches known landmark energy values in the pure spherical spin glasses. In particular, the limit $\alpha\to0$ of the scaled satisfiability transition $V_\text{\textsc{sat}}\sqrt\alpha$ approaches the ground state energy $E_\text{gs}$, while the limit $\alpha\to0$ of the scaled shattering transition $V_\text{sh}\sqrt\alpha$ approaches the threshold energy $E_\text{th}$. The correspondence between ground state and satisfiability is expected: when the energy of a level set is greater in magnitude than the ground state, the level set will usually be empty. The correspondence between the threshold and shattering energies is also intuitive, since the threshold energy is typically understood as the point where the landscape fractures into pieces. However, this second correspondence is only true for the pure spherical models with homogeneous $f(q)$. For any other model with an inhomogeneous $f(q)$, $E_\text{sh}^2V_\text{\textsc{rsb}}^2 \equiv\frac{[f(1)-f(0)]^2}{\alpha f''(0)} -f(0)-\frac{f'(0)}{f''(0)} \end{equation} We conjecture that the \textsc{rsb} instability found in \cite{Urbani_2023_A} is a trait of the cost function \eqref{eq:cost}, and is not inherent to the structure of the solution manifold. Perhaps the best evidence for this is to consider the limit of $M=1$, or $\alpha\to0$ with $E=V_0\sqrt\alpha$ held fixed, where this problem reduces to the level sets of the spherical spin glasses. The instability \eqref{eq:vrsb} implies for the pure spherical 2-spin model with $f(q)=\frac12q^2$ that $E_\textsc{rsb}=\frac12$, though nothing of note is known to occur in the level sets of 2-spin model at such an energy. \section{The quenched shattering energy} \label{sec:1frsb} Here we share how the quenched shattering energy is calculated under a {\oldstylenums1}\textsc{frsb} ansatz. To best make contact with prior work on the spherical spin glasses, we start with \eqref{eq:χ.post-average}. The formula in a quenched calculation is almost the same as that for the annealed, but the order parameters $C$, $R$, $D$, and $G$ must be understood as $n\times n$ matrices rather than scalars. In principle $m$, $\hat m$, $\omega_0$, $\hat\omega_0$, $\omega_1$, and $\hat\omega_1$ should be considered $n$-dimensional vectors, but since in our ansatz replica vectors are constant we can take them to be constant from the start. Expanding the superspace notation, setting $V_0=E\sqrt{N/M}$, and taking $M=1$, we have \begin{align} &\overline{\log\chi(\Omega)} =\lim_{n\to0}\frac\partial{\partial n}\int dC\,dR\,dD\,dG\,dm\,d\hat m\,d\omega_0\,d\hat\omega_0\,d\omega_1\,d\hat\omega_1\, \exp N\Bigg\{ n\hat m +\frac i2\hat\omega_0\operatorname{Tr}(C-I) \notag \\ &\hspace{2em}-\omega_0\operatorname{Tr}(G+R) -in\hat\omega_1E +\frac12\log\det\begin{bmatrix} C-m^2 & i(R-m\hat m) \\ i(R-m\hat m) & D-\hat m^2 \end{bmatrix} -\frac12\log G^2 \notag \\ &\hspace{2em}-\frac12\sum_{ab}^n\left[ \hat\omega_1^2f(C_{ab}) +(2i\omega_1\hat\omega_1R_{ab}+\omega_1^2D_{ab})f'(C_{ab}) +\omega_1^2(G_{ab}^2-R_{ab}^2)f''(C_{ab}) \right] \Bigg\} \end{align} We now make a series of simplifications. Ward identities associated with the BRST symmetry possessed by the original action \cite{Annibale_2003_The, Annibale_2003_Supersymmetric, Annibale_2004_Coexistence} indicate that \begin{align} \omega_1D=-i\hat\omega_1R && G=-R && \hat m=0 \end{align} Moreover, this problem with $m=0$ has a close resemblance to the complexity of the spherical spin glasses. In both, at the BRST-symmetric saddle point the matrix $R$ is diagonal with $R=r_dI$ \cite{Kent-Dobias_2023_How}. To investigate the shattering energy, we can restrict to solutions with $m=0$ and look for the place where such solutions become complex. Inserting these simplifications, we have up to highest order in $N$ \begin{equation} \begin{aligned} \overline{\log\chi(\Omega)} =\lim_{n\to0}\frac\partial{\partial n}\int dC\,dr_d\,d\hat\omega_0\,d\hat\omega_1\, \exp N\Bigg\{ \frac i2\hat\omega_0\operatorname{Tr}(C-I) -in\hat\omega_1E \qquad\\ -i\frac12n\omega_1^*\hat\omega_1r_df'(1) -\frac12\sum_{ab}^n \hat\omega_1^2f(C_{ab}) +\frac12\log\det \left(\frac{-i\hat\omega_1}{\omega_1^*r_d}C+I\right) \Bigg\} \end{aligned} \end{equation} where $\omega_1^*$ is a constant set by satisfying the extremal equations for $D$. If we redefine $\hat\beta=-i\hat\omega_1$ and $\tilde r_d=\omega_1^*r_d$, we find \begin{equation} \begin{aligned} \overline{\log\chi(\Omega)} =\lim_{n\to0}\frac\partial{\partial n}\int dC\,d\hat\beta\,d\tilde r_d\,\hat\omega_0\, \exp N\Bigg\{ \frac i2\hat\omega_0\operatorname{Tr}(C-I) +n\hat\beta E \qquad\\ +n\frac12\hat\beta\tilde r_df'(1) +\frac12\sum_{ab}^n \hat\beta^2f(C_{ab}) +\frac12\log\det \left(\frac{\hat\beta}{\tilde r_d}C+I\right) \Bigg\} \end{aligned} \end{equation} which is exactly the effective action for the supersymmetric complexity in the spherical spin glasses when in the regime where minima dominate \cite{Kent-Dobias_2023_How}. As the effective action for the Euler characteristic, this expression is always valid. Following the same steps as in \cite{Kent-Dobias_2023_How}, we can write the continuum version of this action for arbitrary \textsc{rsb} structure in the matrix $C$ as \begin{equation} \label{eq:cont.action} \frac1N\overline{\log\chi(\Omega)}=\hat\beta E+\frac12\hat\beta\tilde r_df'(1) +\frac12\int_0^1dq\,\left[ \hat\beta^2f''(q)\chi(q)+\frac1{\chi(q)+\tilde r_d\hat\beta^{-1}} \right] \end{equation} where $\chi(q)=\int_1^qdq'\int_0^{q'}dq''P(q'')$ and $P(q)$ is the distribution of off-diagonal elements of the matrix $C$ \cite{Crisanti_1992_The, Crisanti_2004_Spherical, Crisanti_2006_Spherical}. This action must be extremized over the function $\chi$ and the variables $\hat\beta$ and $\tilde r_d$, under the constraint that $\chi(q)$ is continuous, and that it has $\chi'(1)=-1$ and $\chi(1)=0$, necessary for $P$ to be a well-defined probability distribution. Now the specific form of replica symmetry breaking we expect to see is important. We want to study the mixed $2+s$ models in the regime where they may have 1-full \textsc{rsb} in equilibrium \cite{Auffinger_2022_The}. For the Euler characteristic like the complexity, this will correspond to full \textsc{rsb}, in an analogous way to {\oldstylenums1}\textsc{rsb} equilibria give a \textsc{rs} complexity. Such order is characterized by a piecewise smooth $\chi$ of the form \begin{equation} \chi(q)=\begin{cases} \chi_0(q) & q < q_0 \\ 1-q & q \geq q_0 \end{cases} \end{equation} where \begin{equation} \chi_0(q)=\frac1{\hat\beta}[f''(q)^{-1/2}-\tilde r_d] \end{equation} is the function implied by extremizing \eqref{eq:cont.action} over $\chi$ ignoring the continuity and other constraints. The variable $q_0$ must be chosen so that $\chi$ is continuous. The key difference between \textsc{frsb} and {\oldstylenums1}\textsc{frsb} in this setting is that in the former case the ground state has $q_0=1$, while in the latter the ground state has $q_0<1$. \begin{figure} \centering \includegraphics{figs/rsb_comp.pdf} \caption{ \textbf{Self-consistency between \textsc{rsb} instabilities.} Comparison between the predicted value $q_0$ for the \textsc{frsb} solution at the shattering energy in $2+s$ models and the value of the determinant \eqref{eq:stab.det} used in the previous appendix to predict the point of \textsc{rsb} instability. The value of $s$ at which $q_0$ becomes nonzero is precisely the point where the determinant has a nontrivial zero. } \label{fig:rsb} \end{figure} We use this action to find the shattering energy in the following way. First, we know that the ground state energy is the place where the manifold and therefore the average Euler characteristic vanishes. Therefore, setting $\overline{\log\chi(\Omega)}=0$ and solving for $E$ yields a formula for the ground state energy \begin{equation} E_\text{gs}=-\frac1{\hat\beta}\left\{ \frac12\hat\beta\tilde r_df'(1) +\frac12\int_0^1dq\,\left[ \hat\beta^2f''(q)\chi(q)+\frac1{\chi(q)+\tilde r_d\hat\beta^{-1}} \right] \right\} \end{equation} This expression can be maximized over $\hat\beta$ and $\tilde r_d$ to find the correct parameters at the ground state for a particular model. Then, the shattering energy is found by slowly lowering $q_0$ and solving the combined extremal and continuity problem for $\hat\beta$, $\tilde r_d$, and $E$ until $E$ reaches a maximum value and starts to decrease. This maximum is the shattering energy, since it is the point where the $m=0$ solution becomes complex. Starting from this point, we take small steps in $s$ and $\lambda_s$, simultaneously extremizing, ensuring continuity, and maximizing $E$. This draws out the shattering energy across the entire range of $s$ plotted in Fig.~\ref{fig:ssg}. The transition to the \textsc{rs} solution occurs when the value $q_0$ that maximizes $E$ hits zero. We find that the transition between \textsc{rs} and \textsc{frsb} is precisely predicted by the \textsc{rsb} instability calculated in Appendix~\ref{sec:rms}, as shown in Fig.~\ref{fig:rsb}. \bibliography{topology} \end{document}