\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} \usepackage[dvipsnames]{xcolor} \usepackage{anyfontsize,authblk} \usepackage{xfrac} % sfrac \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} } \author{\footnote{\url{jaron.kent-dobias@roma1.infn.it}}} \affil{Istituto Nazionale di Fisica Nucleare, Sezione di Roma I, Italy} \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$^\star$} \end{center} \begin{center} Istituto Nazionale di Fisica Nucleare, Sezione di Roma I, Italy \\[\baselineskip] $\star$ \href{mailto:jaron.kent-dobias@roma1.infn.it}{\small jaron.kent-dobias@roma1.infn.it} \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 polynomial coefficients. We divide the behavior in four phases: a connected phase, an onset phase, a shattered phase, and an \textsc{unsat} phase. In the connected phase, the average characteristic is 2 and there is a single extensive connected component, while in the onset phase the Euler characteristic is exponentially large in $N$. In the shattered phase the characteristic remains exponentially large but subextensive components appear, while in the \textsc{unsat} phase the manifold vanishes. 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 between the onset and shattered phases corresponds to the asymptotic limit of gradient descent from a random initial condition. } } \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 such problems, there is great interest in characterizing structure in the set of solutions, which can be influential in how algorithms behave when trying to solve them \cite{Baldassi_2016_Unreasonable, Baldassi_2019_Properties, Beneventano_2023_On}. Here, we show how topological information about the set of solutions can be calculated in a simple model of satisfying random nonlinear equalities. This allows us to reason about the connectivity of this 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 in gradient descent 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 $V_k$. In particular, 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} with the elements of the tensors $J^{(k,p)}$ as independently distributed unit normal random variables satisfies \eqref{eq:covariance}. The size of the series coefficients of $f$ therefore control the variances in the coefficients of random polynomial constraints. When $M=1$, this problem corresponds to the level set of a spherical spin glass with 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. 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 two related means: its average Euler characteristic, and the average number of stationary points of a linear height function restricted to the manifold. These measures tell us complementary pieces of information. 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{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 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 to compute $\chi$ 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 \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 $H$ 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. To evaluate the average of $\chi$ over the constraints, we first translate the $\delta$ functions and determinant to integral form, with \begin{align} \label{eq:delta.exp} \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)} \\ \label{eq:det.exp} \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} \label{eq:pre-saddle.characteristic} \overline{\chi(\Omega)} =\left(\frac N{2\pi}\right)^2\int dR\,dD\,dm\,d\hat m\,g(R,D,m,\hat m)\,e^{N\mathcal S_\chi(R,D,m,\hat m)} \end{equation} where $g$ is a prefactor of $o(N^0)$, and $\mathcal S_\chi$ is an effective action defined by \begin{equation} \label{eq:euler.action} \begin{aligned} \mathcal S_\chi(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 to leading order 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^*}{1-m^2}R^* \qquad \hat m=0 \end{equation} \begin{equation} \begin{aligned} R^* =\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} with the resulting effective action as a function of $m$ alone given by \begin{equation} \mathcal S_\chi(m) =-\frac\alpha2\bigg[ \log\left( 1-\frac{f(1)}{f'(1)}\frac{1+\frac m{R^*}}{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^*}} \right)^{-1} \bigg] +\frac12\log\left(-\frac m{R^*}\right) \end{equation} This function is plotted as a function of $m$ in Fig.~\ref{fig:action} for a variety of $V_0$ and $f$. 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. \begin{figure} \includegraphics{figs/action_1.pdf} \hspace{-3.5em} \includegraphics{figs/action_3.pdf} \caption{ \textbf{Effective action for the Euler characteristic.} The effective action governing the average Euler characteristic as a function of the overlap $m=\frac1N\mathbf x\cdot\mathbf x_0$ with the height direction for two different homogeneous polynomial functions and a variety of target values $V_0$. 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 after the satisfiability transition at $V_0=V_\text{\textsc{sat}}=1$, $m^*$ goes to zero and the action becomes negative. \textbf{Right:} With nonlinear functions there are four regimes. For small $V_0$ the behavior is the same as in the linear case, with zero action. After an onset transition at $V_0=V_\text{on}\simeq1.099$ the maxima are at the edge of validity of the action and the action is positive. At a shattering transition at $V_0=V_\text{sh}\simeq1.394$, the maximum goes to zero and the action is positive. Finally, at the satisfiability transition $V_0=V_\text{\textsc{sat}}\simeq1.440$ the action becomes negative. } \label{fig:action} \end{figure} The action $\mathcal S_\chi$ is extremized with respect to $m$ at $m=0$ or $m=\pm m^*=\mp R^*(m^*)$. In the latter case, $m^*$ takes the value \begin{equation} m^*=\sqrt{1-\frac{\alpha}{f'(1)}\big(V_0^2+f(1)\big)} \end{equation} and $\mathcal S_\chi(m^*)=0$. 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 becomes complex in the region $m^20$, the solution at $m=0$ is also not valid. In fact, it is not clear what the average value of the Euler characteristic should be at all when there is some range $-m_\text{min}0$ and $\operatorname{Re}\mathcal S_\chi(0)<0$. Here, $\overline{\chi(\Omega)}=2+o(1)$ for even $N-M-1$, strongly indicating a topology homeomorphic to the $S^{N-M-1}$ sphere. \item \textbf{Regime II: \boldmath{$\overline{\chi(\Omega)}$} very large and negative, isolated contribution at \boldmath{$m=\pm m^*$}.} Here $\frac1N\log\overline{\chi(\Omega)^2}>0$ and $\overline{\chi(\Omega)}<0$. This regime occurs when $m_\text{min}^2>0$ and $\operatorname{Re}\mathcal S_\chi(0)>0$. Here the average Euler characteristic is large and negative. While the topology of the manifold is not necessarily connected in this regime, holes are more numerous than components. In addition, here $m_*^2>m_\text{min}^2$, meaning that the solutions at $m=\pm m_*$ are still present. This indicates that the manifold has topologically simple boundaries with some separation from the sea of holes. \item \textbf{Regime III: \boldmath{$\overline{\chi(\Omega)}$} very large and negative, no contribution at \boldmath{$m=\pm m^*$}.} The same as Regime II, but with $m_*^20$ and $\overline{\chi(\Omega)}>0$. This regime occurs when $m_\text{min}^2<0$ and $\mathcal S_\chi(0)>0$. Here the average Euler characteristic is large and positive. Large connected components of the manifold may or may not exist, but small disconnected components outnumber holes. \item \textbf{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. \end{itemize} \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 model for three different homogeneous covariance functions. The onset transition $V_\text{on}$, shattering transition $V_\text{sh}$, and satisfiability transition $V_\text{\textsc{sat}}$ are indicated when they exist. In the limit of $\alpha\to0$, the behavior of level sets of the spherical spin glasses are recovered: the final plot shows how in the pure cubic model the ground state energy $E_\text{gs}$ and threshold energy $E_\text{th}$ correspond with the limits of the satisfiability and shattering transitions, respectively. 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} However, when the magnitude of $V_0$ is sufficiently large, with \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 this solution is no longer valid. Since \begin{equation} V_\text{on}^2-V_{\text{\textsc{sat}}\ast}^2 =\frac{f'(1)-f(1)}\alpha-\frac{\sqrt{1-\alpha}}\alpha f(1) \end{equation} If $f(q)$ is purely linear, then $f'(1)=f(1)$ and $V_\text{on}^2>V_{\text{\textsc{sat}}\ast}^2$, so the naive 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 $f$. Likewise, the solution at $m=0$ is not always valid. For $V_0$ of magnitude sufficiently small, with \begin{equation} V_0^2V_\text{\textsc{rsb}}^2 \equiv\frac{[f(1)-f(0)]^2}{\alpha f''(0)} -f(0)-\frac{f'(0)}{f''(0)} \end{equation} Some preliminary attempts to find an instability towards finite-\textsc{rsb} in this region in the calculation of the Euler characteristic did not turn up anything. 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{Implications for the dynamics of spherical spin glasses} \label{sec:ssg} As indicated earlier, for $M=1$ the solution manifold corresponds to the energy level set of a spherical spin glass with energy density $E=\sqrt NV_0$. All the results from the previous sections follow, and can be translated to the spin glasses by taking the limit $\alpha\to0$ while keeping $E=V_0\alpha^{-1/2}$ fixed. With a little algebra this procedure yields \begin{align} E_\text{on}=\pm\sqrt{2f(1)} && E_\text{sh}=\pm\sqrt{4f(1)\left(1-\frac{f(1)}{f'(1)}\right)} \label{eq:ssg.energies} \end{align} for the energies at which level sets of the spherical spin glasses have disconnected pieces appear, and that at which a large connected component vanishes. For the pure $p$-spin spherical spin glasses with $f(q)=\frac12q^p$, $E_\text{sh}=\sqrt{2(p-1)/p}$, precisely the threshold energy in these models \cite{Castellani_2005_Spin-glass}. This is expected, since threshold energy, defined as the place where marginal minima are dominant in the landscape, is widely understood as the place where level sets are broken into pieces. However, for general mixed models the threshold energy is \begin{equation} E_\mathrm{th}=\pm\frac{f(1)[f''(1)-f'(1)]+f'(1)^2}{f'(1)\sqrt{f''(1)}} \end{equation} which satisfies $|E_\text{sh}|\leq|E_\text{th}|$. Therefore, as one descends in energy in generic models one will meet the shattering energy before the threshold energy. This is perhaps unexpected, since one wight imagine that where level sets of the energy break into many pieces would coincide with the largest concentration of shallow minima in the landscape. We see here that this isn't the case. This fact mirrors another another that was made clear recently: when gradient decent dynamics are run on these models, they will asymptotically reach an energy above the threshold energy \cite{Folena_2020_Rethinking, Folena_2021_Gradient, Folena_2023_On}. The old belief that the threshold energy qualitatively coincides with a kind of shattering is the origin of the expectation that the dynamic limit should be the threshold energy. Given that our work shows the actual shattering energy is different, here we compare it with data on the asymptotic limits of dynamics to see if they coincide. \begin{figure} \includegraphics{figs/dynamics_2.pdf} \hspace{-0.5em} \includegraphics{figs/dynamics_3.pdf} \caption{ \textbf{Is the shattering energy a dynamic threshold?} Comparison of the shattering energy $E_\text{sh}$ with the asymptotic performance of gradient descent from a random initial condition in $p+s$ models with $p=2$ and $p=3$ and varying $s$. The values of $\lambda$ depend on $p$ and $s$ and are taken from \cite{Folena_2023_On}. The points show the asymptotic performance extrapolated using two different methods and have unknown uncertainty, from \cite{Folena_2023_On}. Also shown is the annealed threshold energy $E_\text{th}$, where marginal minima are the most common type of stationary point. The section of $E_\text{sh}$ that is dashed on the left plot indicates the continuation of the annealed result, whereas the solid portion gives the value calculated with a {\oldstylenums 1}\textsc{frsb} ansatz. } \label{fig:ssg} \end{figure} Measurements of the asymptotic limits of dynamics were recently taken in \cite{Folena_2023_On} for two different classes of models with inhomogeneous $f(q)$, with \begin{equation} f(q)=\frac12\big[\lambda q^p+(1-\lambda)q^s\big] \end{equation} They studied models with this covariance for $p=2$ and $p=3$ while varying $s$. In both cases, the relative weight $\lambda$ varies with $s$ and was chosen to maximize a heuristic to increase the chances of seeing nontrivial behavior. The authors numerically integrated the dynamic mean field theory (\textsc{dmft}) equations for gradient descent on these models from a random initial condition to large but finite time, then attempted to extrapolate the infinite-time behavior by two different methods. The black symbols in Fig.~\ref{fig:ssg} show the measurements taken from \cite{Folena_2023_On}. The difference between the two extrapolations is not critical here, see the original paper for details. We simply note that the authors of \cite{Folena_2023_On} did not associate an uncertainty with them, nor were they confident that they are unbiased estimates of the asymptotic value. Fig.~\ref{fig:ssg} also shows the shattering and {\oldstylenums1}\textsc{rsb} threshold energies as a function of $s$. The solid lines come from using \textsl{Mathematica}'s \texttt{Interpolation} function to create a smooth function $\lambda(s)$ through the values used in \cite{Folena_2023_On}. For the $2+s$ models for sufficiently large $s$, the ground state is described by a {\oldstylenums1}\textsc{frsb} order and both the threshold energy and shattering energy calculated using an annealed average are likely inaccurate \cite{Auffinger_2022_The}. In Appendix~\ref{sec:1frsb} we calculate the quenched ground state and shattering energies for these models consistent with the {\oldstylenums1}\textsc{frsb} equilibrium order. In the left panel of Fig.~\ref{fig:ssg}, the solid line shows the quenched calculation, while the dashed line shows the annealed formula \eqref{eq:ssg.energies}. Is the shattering energy consistent with the dynamic threshold for gradient descent from a random initial condition? The evidence in Fig.~\ref{fig:ssg} is compelling but inconclusive. The difference between the shattering energy and the extrapolated \textsc{dmft} data is about the same as the difference between the values predicted by the two extrapolation methods. If both extrapolation methods suffer from similar systematic biases, it is plausible they true value corresponds with the shattering energy. However, better estimates of the asymptotic values are needed to support or refute this conjecture. This motivates working to integrate the \textsc{dmft} equations to longer times, or else look for analytic asymptotic solutions that approach $E_\text{sh}$. \section{Conclusion} We have shown how to calculate the average Euler characteristic of the solution manifold in a simple model of random constraint satisfaction. The results constrain the topology and geometry of this manifold, revealing when it is connected and trivial, when it is extensive but topologically nontrivial, and when it is shattered into disconnected pieces. These inferences are supported by a auxiliary calculation of the complexity of stationary points for a test function on the same solution manifold. This calculation has novel implications for the geometry of the energy landscape in the spherical spin glasses, where it reveals a previously unknown landmark energy $E_\text{sh}$. This shattering energy is where the topological calculation implies that the level set of the energy breaks into disconnected pieces, and differs from the threshold energy $E_\text{th}$ in mixed models. It's possible that $E_\text{sh}$ is the asymptotic limit of gradient descent from a random initial condition in such models, but the quality of the currently available data makes this conjecture inconclusive. This paper has focused on equality constraints, while most existing studies of constraint satisfaction study inequality constraints \cite{Franz_2016_The, Franz_2017_Universality, Franz_2019_Critical, Annesi_2023_Star-shaped, Baldassi_2023_Typical}. To generalize the technique developed in this paper to such cases is not a trivial extension. The set of solutions to such problems are manifolds with boundary, and these boundaries are often not smooth. To study such cases with these techniques will require using extensions of the Morse theory for manifolds with boundary, and will be the subject of future work. \paragraph{Acknowledgements} The authors thank Pierfrancesco Urbani for helpful conversations on these topics, and Giampaolo Folena for supplying his \textsc{dmft} data for the spherical spin glasses. \paragraph{Funding information} JK-D is supported by a \textsc{DynSysMath} Specific Initiative of the INFN. \appendix \section{Details of the calculation of the average Euler characteristic} \label{sec:euler} Our starting point is the expression \eqref{eq:kac-rice.lagrange} with the substitutions of the $\delta$-function and determinant \eqref{eq:delta.exp} and \eqref{eq:det.exp} made. To make the calculation compact, we introduce superspace coordinates \cite{DeWitt_1992_Supermanifolds}. Introducing the Grassmann indices $\bar\theta_1$ and $\theta_1$, we define the supervectors \begin{align} \pmb\phi(1)=\mathbf x+\bar\theta_1\pmb\eta+\bar{\pmb\eta}\theta_1+\bar\theta_1\theta_1i\hat{\mathbf x} && \sigma_k(1)=\omega_k+\bar\theta_1\gamma_k+\bar\gamma_k\theta_1+\bar\theta_1\theta_1i\hat\omega_k \end{align} with associated measures \begin{align} d\pmb\phi=d\mathbf x\,\frac{d\hat{\mathbf x}}{(2\pi)^N}\,d\bar{\pmb\eta}\,d\pmb\eta && d\pmb\sigma=d\pmb\omega\,\frac{d\hat{\pmb\omega}}{(2\pi)^{M+1}}\,d\bar{\pmb\gamma}\,d\pmb\gamma \end{align} The Euler characteristic can be expressed using these supervectors as \begin{align} &\chi(\Omega) =\int d\pmb\phi\,d\pmb\sigma\,e^{\int d1\,L(\pmb\phi(1),\pmb\sigma(1))} \\ &=\int d\pmb\phi\,d\pmb\sigma\,\exp\left\{ \int d1\left[ H\big(\pmb\phi(1)\big) +\frac12\sigma_0(1)\left(\|\pmb\phi(1)\|^2-N\right) +\sum_{k=1}^M\sigma_k(1)\Big(V_k\big(\pmb\phi(1)\big)-V_0\Big) \right] \right\} \notag \end{align} where $d1=d\bar\theta_1\,d\theta_1$ is the integral over both Grassmann indices. Since this is an exponential integrand linear in the Gaussian functions $V_k$, we can take their average to find \begin{equation} \label{eq:χ.post-average} \begin{aligned} \overline{\chi(\Omega)} =\int d\pmb\phi\,d\pmb\sigma\,\exp\Bigg\{ \int d1\left[ H(\pmb\phi(1)) +\frac12\sigma_0(1)\big(\|\pmb\phi(1)\|^2-N\big) -V_0\sum_{k=1}^M\sigma_k(1) \right] \\ +\frac12\int d1\,d2\,\sum_{k=1}^M\sigma_k(1)\sigma_k(2)f\left(\frac{\pmb\phi(1)\cdot\pmb\phi(2)}N\right) \Bigg\} \end{aligned} \end{equation} This is a super-Gaussian integral in the super-Lagrange multipliers $\sigma_k$ with $1\leq k\leq M$. Performing that integral yields \begin{equation} \begin{aligned} \overline{\chi(\Omega)} &=\int d\pmb\phi\,d\sigma_0\,\exp\Bigg\{ \int d1\left[ H(\pmb\phi(1)) +\frac12\sigma_0(1)\big(\|\pmb\phi(1)\|^2-N\big) \right] \\ &\hspace{5em}-\frac M2V_0^2\int d1\,d2\,f\left(\frac{\pmb\phi(1)\cdot\pmb\phi(2)}N\right)^{-1} -\frac M2\log\operatorname{sdet}f\left(\frac{\pmb\phi(1)\cdot\pmb\phi(2)}N\right) \Bigg\} \end{aligned} \end{equation} The supervector $\pmb\phi$ enters this expression as a function only of the scalar product with itself and with the vector $\mathbf x_0$ inside the height function $H(\mathbf x)=\mathbf x_0\cdot\mathbf x$. We therefore make a change of variables to the superoperator $\mathbb Q$ and the supervector $\mathbb M$ defined by \begin{equation} \mathbb Q(1,2)=\frac{\pmb\phi(1)\cdot\pmb\phi(2)}N \qquad \mathbb M(1)=\frac{\pmb\phi(1)\cdot\mathbf x_0}N \end{equation} These new variables can replace $\pmb\phi$ in the integral using a generalized Hubbard--Stratonovich transformation, which yields \begin{align} &\overline{\chi(\Omega)} =\frac12\int d\mathbb Q\,d\mathbb M\,d\sigma_0\, \left([\operatorname{sdet}(\mathbb Q-\mathbb M\mathbb M^T)]^\frac12+O(N^{-1})\right) \,\exp\Bigg\{ \frac N2\log\operatorname{sdet}(\mathbb Q-\mathbb M\mathbb M^T) \label{eq:post.hubbard-strat} \\ &+N\int d1\left[ \mathbb M(1) +\frac12\sigma_0(1)\big(\mathbb Q(1,1)-1\big) \right] -\frac M2V_0^2\int d1\,d2\,f(\mathbb Q)^{-1}(1,2) -\frac M2\log\operatorname{sdet}f(\mathbb Q) \Bigg\} \notag \end{align} where we show the asymptotic value of the prefactor in Appendix~\ref{sec:prefactor}. To move on from this expression, we need to expand the superspace notation. We can write \begin{equation} \label{eq:ops.q} \begin{aligned} \mathbb Q(1,2) &=C-R(\bar\theta_1\theta_1+\bar\theta_2\theta_2) -G(\bar\theta_1\theta_2+\bar\theta_2\theta_1) -D\bar\theta_1\theta_1\bar\theta_2\theta_2 \\ &\qquad +(\bar\theta_1+\bar\theta_2)H +\bar H(\theta_1+\theta_2) -(\bar\theta_1\theta_1\bar\theta_2+\bar\theta_2\theta_2\bar\theta_1)i\hat H -\bar{\hat H}(\theta_1\bar\theta_2\theta_2+\theta_1\bar\theta_1\theta_1) \end{aligned} \end{equation} and \begin{equation} \mathbb M(1) =m+\bar\theta_1H_0+\bar H_0\theta_1+i\hat m\bar\theta_1\theta_1 \label{eq:ops.m} \end{equation} with associated measures \begin{align} d\mathbb Q =dC\,dR\,dG\,\frac{dD}{(2\pi)^2}\,d\bar H\,dH\,d\bar{\hat H}\,d\hat H && d\mathbb M=dm\,\frac{d\hat m}{2\pi}\,d\bar H_0\,dH_0 \label{eq:op.measures} \end{align} The order parameters $C$, $R$, $G$, $D$, $m$, and $\hat m$ are ordinary numbers defined by \begin{align} \label{eq:ops.bos} C=\frac{\mathbf x\cdot\mathbf x}N && R=-i\frac{\mathbf x\cdot\hat{\mathbf x}}N && G=\frac{\bar{\pmb\eta}\cdot\pmb\eta}N && D=\frac{\hat{\mathbf x}\cdot\hat{\mathbf x}}N && m=\frac{\mathbf x_0\cdot\mathbf x}N && \hat m=-i\frac{\mathbf x_0\cdot\hat{\mathbf x}}N \end{align} while $\bar H$, $H$, $\bar{\hat H}$, $\hat H$, $\bar H_0$ and $H_0$ are Grassmann numbers defined by \begin{align} \label{eq:ops.ferm} \bar H=\frac{\bar{\pmb\eta}\cdot\mathbf x}N && H=\frac{\pmb\eta\cdot\mathbf x}N && \bar{\hat H}=-i\frac{\bar{\pmb\eta}\cdot\hat{\mathbf x}}N && \hat H=-i\frac{\pmb\eta\cdot\hat{\mathbf x}}N && \bar H_0=\frac{\bar{\pmb\eta}\cdot\mathbf x_0}N && H_0=\frac{\pmb\eta\cdot\mathbf x_0}N \end{align} We can treat the integral over $\sigma_0$ immediately. It gives \begin{equation} \label{eq:sigma0.integral} \int d\sigma_0\,e^{N\int d1\,\frac12\sigma_0(1)(\mathbb Q(1,1)-1)} =2\times2\pi\delta(C-1)\,\delta(G+R)\,\bar HH \end{equation} This therefore sets $C=1$ and $G=-R$ in the remainder of the integrand, as well as removing all dependence on $\bar H$ and $H$. With these solutions inserted, the remaining terms in the exponential give \begin{align} \label{eq:sdet.q} &\operatorname{sdet}(\mathbb Q-\mathbb M\mathbb M^T) =1+\frac{(1-m^2)D+\hat m^2-2Rm\hat m}{R^2} -\frac6{R^4}\bar H_0H_0\bar{\hat H}\hat H \\ &\hspace{5pc}+\frac2{R^3}\left[ (mR-\hat m)(\bar{\hat H}H_0+\bar H_0\hat H) -(D+R^2)\bar H_0H_0 +(1-m^2)\bar{\hat H}\hat H \right] \notag \\ \label{eq:sdet.fq} &\operatorname{sdet}f(\mathbb Q) =1+\frac{Df(1)}{R^2f'(1)} +\frac{2f(1)}{R^3f'(1)}\bar{\hat H}\hat H \\ \label{eq:inv.q} &\int d1\,d2\,f(\mathbb Q)^{-1}(1,2) =\left(f(1)+\frac{R^2f'(1)}{D}\right)^{-1} +2\frac{Rf'(1)}{(Df(1)+R^2f'(1))^2}\bar{\hat H}\hat H \end{align} The Grassmann terms in these expressions do not contribute to the effective action, but will be important in our derivation of the prefactor for the calculation in the connected regime. The substitution of these expressions into \eqref{eq:post.hubbard-strat} without the Grassmann terms yields \eqref{eq:euler.action} from the main text. \section{Calculation of the prefactor of the average Euler characteristic} \label{sec:prefactor} Because of our convention of including the appropriate factors of $2\pi$ in the superspace measure, super-Gaussian integrals do not produce such factors in our derivation. Prefactors to our calculation come from three sources: the introduction of $\delta$-functions to define the order parameters, integrals over Grassmann order parameters, and from the saddle point approximation to the large-$N$ integral. In addition, there are important contributions of a sign of the magnetization at the solution that arise from our super-Gaussian integrations. \subsection{Contribution from the Hubbard--Stratonovich transformations} \label{sec:prefactor.hs} First, we examine the factors arising from the definition of order parameters. This begins by introducing to the integral the factor of one \begin{equation} 1=(2\pi)^3\int d\mathbb Q\,d\mathbb M\, \delta\big(N\mathbb Q(1,2)-\pmb\phi(1)\cdot\pmb\phi(2)\big)\, \delta\big(N\mathbb M(1)-\mathbf x_0\cdot\pmb\phi(1)\big) \end{equation} where three factors of $2\pi$ come from the measures as defined in \eqref{eq:op.measures}. Converting the $\delta$-function into an exponential integral yields \begin{equation} \begin{aligned} 1=\frac12\int d\mathbb Q\,d\mathbb M\,d\tilde{\mathbb Q}\,d\tilde{\mathbb M}\, \exp\Bigg\{ \frac12\int d1\,d2\,\tilde{\mathbb Q}(1,2)\big(N\mathbb Q(1,2)-\pmb\phi(1)\cdot\pmb\phi(2)\big) \qquad \\ +\int d1\,i\tilde{\mathbb M}(1)\big(N\mathbb M(1)-\mathbf x_0\cdot\pmb\phi(1)\big) \Bigg\} \end{aligned} \end{equation} where the supervectors and measures for $\tilde{\mathbb Q}$ and $\tilde{\mathbb M}$ are defined analogously to those of $\mathbb Q$ and $\mathbb M$. This is now a super-Gaussian integral in $\pmb\phi$, which can be performed to yield \begin{equation} \begin{aligned} \int d\pmb\phi\,1=\frac12\int d\mathbb Q\,d\mathbb M\,d\tilde{\mathbb Q}\,d\tilde{\mathbb M}\, \exp\Bigg\{ \frac N2\int d1\,d2\,\tilde{\mathbb Q}(1,2)\mathbb Q(1,2) +N\int d1\,i\tilde{\mathbb M}(1)\mathbb M(1) \quad \\ -\frac N2\log\operatorname{sdet}\tilde{\mathbb Q} -\frac N2\int d1\,d2\,\tilde{\mathbb M}(1)\tilde{\mathbb Q}^{-1}(1,2)\tilde{\mathbb M}(2) \Bigg\} \end{aligned} \end{equation} We can perform the remaining Gaussian integral in $\tilde{\mathbb M}$ to find \begin{equation} \begin{aligned} \int d\pmb\phi\,1= \frac12\int d\mathbb Q\,d\mathbb M\,d\tilde{\mathbb Q}\, (\operatorname{sdet}\tilde{\mathbb Q}^{-1})^{-\frac12} \exp\Bigg\{ -\frac N2\log\operatorname{sdet}\tilde{\mathbb Q} \hspace{5em} \\ \frac N2\int d1\,d2\,\tilde{\mathbb Q}(1,2)\big[ \mathbb Q(1,2)-\mathbb M(1)\mathbb M(2) \big] \Bigg\} \end{aligned} \end{equation} The integral over $\tilde{\mathbb Q}$ can be evaluated to leading order using the saddle point method. The integrand is stationary at $\tilde{\mathbb Q}=(\mathbb Q-\mathbb M\mathbb M^T)^{-1}$, and substituting this into the above expression results in the term $\frac12\log\det(\mathbb Q-\mathbb M\mathbb M^T)$ in the effective action from \eqref{eq:post.hubbard-strat}. The saddle point also yields a prefactor of the form \begin{equation} \left(\operatorname{sdet}_{\{1,2\},\{3,4\}}\frac{\partial^2\frac12\log\operatorname{sdet}\tilde{\mathbb Q}}{\partial\tilde{\mathbb Q}(1,2)\partial\tilde{\mathbb Q}(3,4)}\right)^{-\frac12} =\left(\operatorname{sdet}_{\{1,2\},\{3,4\}}\tilde{\mathbb Q}^{-1}(3,1)\tilde{\mathbb Q}^{-1}(2,4)\right)^{-\frac12} =1 \end{equation} where the final superdeterminant is identically 1 for any superoperator $\tilde{\mathbb Q}$, not just its saddle-point value. The Hubbard--Stratonovich transformation therefore contributes a factor of \begin{equation} \frac12\operatorname{sdet}(\mathbb Q-\mathbb M\mathbb M^T)^{\frac12} =\frac12[(C-m^2)(D+\hat m^2)+(R-m\hat m)^2]^\frac12G^{-1} \end{equation} to the prefactor at the largest order in $N$. \subsection{Sign of the prefactor} The superspace notation papers over some analytic differences between branches of the logarithm that are not important for determining the saddle point but are important to getting correctly the sign of the prefactor. For instance, consider the superdeterminant of $\mathbb Q$ from \eqref{eq:ops.q} (dropping the fermionic order parameters for a moment for brevity), \begin{equation} \operatorname{sdet}\mathbb Q=\frac{CD+R^2}{G^2} \end{equation} The numerator and denominator arise from the determinant in the sector of number and Grassmann number basis elements for the superoperator, respectively. In our calculation, such superdeterminants appear after Gaussian integrals, which produce \begin{equation} \int d\pmb\phi\,\exp\left\{-\frac12\int d1\,d2\,\pmb\phi(1)\mathbb Q(1,2)\pmb\phi(2)\right\} =(\operatorname{sdet}\mathbb Q)^{-\frac12} =(CD+R^2)^{-\frac12}G \end{equation} Here we emphasize that in the expanded result of the integral, the term from the denominator of the square root enters not as $(G^2)^{\frac12}=|G|$ but as $G$, including its sign. Therefore, when we write in the effective action $\frac12\log\operatorname{sdet}\mathbb Q$, we should really be writing \begin{equation} \int d\pmb\phi\,\exp\left\{-\frac12\int d1\,d2\,\pmb\phi(1)\mathbb Q(1,2)\pmb\phi(2)\right\} =\operatorname{sign}(G)e^{-\frac12\log\operatorname{sdet}\mathbb Q} \end{equation} In our calculation in Appendix \ref{sec:euler} we elide this several times, and accumulate $M$ factors of $\operatorname{sign}(-Gf'(C))=\operatorname{sign}(-G)$ from the Gaussian integral over Lagrange multipliers and $N$ factors of $\operatorname{sign}(-G)$ from the Hubbard--Stratonovich transformation. Since at all saddle points $G=-R$, we have \begin{equation} \operatorname{sign}(R)^{N+M}e^{N\mathcal S_\chi(\tilde{\mathbb Q},\mathbb Q,\mathbb M)} \end{equation} \subsection{Contribution from integrating the Grassmann order parameters} After integrating out the Lagrange multiplier enforcing the spherical constraint in \eqref{eq:sigma0.integral}, the Grassmann variables $\bar H$ and $H$ are eliminated from the integrand. This leaves dependence on $\bar{\hat H}$, $\hat H$, $\bar H_0$, and $H_0$. Expanding the contributions from \eqref{eq:sdet.q}, \eqref{eq:sdet.fq}, and \eqref{eq:inv.q}, the contribution to the action is given by \begin{equation} \int d\bar{\hat H}\,d\hat H\,d\bar H_0\,dH_0\,\exp\left\{ N\begin{bmatrix} \bar{\hat H} & \bar H_0 \end{bmatrix} \begin{bmatrix} h_1 & h_2 \\ h_2 & h_3 \end{bmatrix} \begin{bmatrix} \hat H \\ H_0 \end{bmatrix} +Nh_4\bar{\hat H}\hat H\bar H_0H_0 \right\} =N^2(h_1h_3-h_2^2)+Nh_4 \end{equation} where \begin{align} &h_1=\frac1R\left( \frac{1-m^2}{D(1-m^2)+R^2-2Rm\hat m+\hat m^2} -\alpha\frac{Df(1)^2+R^2f'(1)[V_0^2+f(1)]}{[Df(1)+R^2f'(1)]^2} \right) \\ &h_2=\frac1R\frac{Rm-\hat m}{D(1-m^2)+R^2-2Rm\hat m+\hat m^2} \qquad h_3=-\frac1R\frac{D+R^2}{D(1-m^2)+R^2-2Rm\hat m+\hat m^2} \\ &h_4=-\frac1{R^2}\frac1{D(1-m^2)+R^2-2Rm\hat m+\hat m^2} \end{align} The contribution at leading order in $N$ is therefore \begin{equation} \frac{N^2}{R^2[D(1-m^2)+R^2-2Rm\hat m+\hat m^2]}\left( \alpha\frac{(D+R^2)[Df(1)^2+R^2f'(1)[V_0^2+f(1)]]}{ [Df(1)+R^2f'(1)]^2 } -1 \right) \end{equation} \subsection{Contribution from the saddle point approximation} We now want to evaluate the prefactor for the asymptotic value of $\overline{\chi(\Omega)}$. From the previous sections, the definition of the measures $d\mathbb Q$ and $d\mathbb M$ in \eqref{eq:op.measures}, and the integral over $\sigma_0$ of \eqref{eq:sigma0.integral}, we can now see that the function $g(R,D,m,\hat m)$ of \eqref{eq:pre-saddle.characteristic} is given by \begin{equation} \begin{aligned} g(R,D,m,\hat m) =- \frac{\operatorname{sign}(R)^{N+M}}{R^3[D(1-m^2)+R^2-2Rm\hat m+\hat m^2]^{\frac12}} \hspace{3pc} \\ \times\left( \alpha\frac{(D+R^2)[Df(1)^2+R^2f'(1)[V_0^2+f(1)]]}{ [Df(1)+R^2f'(1)]^2 } -1 \right) \end{aligned} \end{equation} In the connected regime, there are two saddle points of the integrand that contribute to the asymptotic value of the integral, at $m=\pm m^*$ with $R=-m^*$, $D=0$, and $\hat m=0$. At this saddle point $\mathcal S_\chi=0$. We can therefore write \begin{equation} \overline{\chi(\Omega)} =\sum_{m=\pm m^*} g(-m,0,m,0)\big[\det\partial\partial\mathcal S_\chi(-m,0,m,0)\big]^{-\frac12} \end{equation} where here $\partial=[\frac{\partial}{\partial R},\frac{\partial}{\partial D},\frac\partial{\partial m},\frac\partial{\partial \hat m}]$ is the vector of derivatives with respect to the remaining order parameters. For both of the two saddle points, the determinant of the Hessian of the effective action evaluates to \begin{equation} \det\partial\partial\mathcal S_\chi =\left[\frac1{(m^*)^4}\left(1-\frac{\alpha[V_0^2+f(1)]}{f'(1)}\right)\right]^2 \end{equation} whereas \begin{equation} g(\mp m^*,0,\pm m^*,0) =\frac{(\mp1)^{N+M+1}}{|m^*|^4}\left(1-\frac{\alpha[V_0^2+f(1)]}{f'(1)}\right) \end{equation} The saddle point at $m=-m^*$, characterized by minima of the height function, always contributes with a positive term. On the other hand, the saddle point with $m=+m^*$, characterized by maxima of the height function, contributes with a sign depending on if $N+M+1$ is even or odd. This follows from the fact that minima, with an index of 0, have a positive contribution to the sum over stationary points, while maxima, with an index of $N-M-1$, have a contribution that depends on the dimension of the manifold. We have finally that, in the connected regime, \begin{equation} \overline{\chi(\Omega)}=1+(-1)^{N+M+1}+O(N^{-1}) \end{equation} When $N+M+1$ is odd, this evaluates to zero. In fact it must be zero to all orders in $N$, since for odd-dimensional manifolds the Euler characteristic is always zero. When $N+M+1$ is even, we have $\overline{\chi(\Omega)}=2$ to leading order in $N$, as specified in the main text. \section{The root mean square Euler characteristic} \label{sec:rms} Here we calculate $\overline{\chi(\Omega)^2}$, the average of the squared Euler characteristic. This is accomplished by taking two copies of the integral, with \begin{equation} \chi(\Omega)^2 =\int d\pmb\phi_1\,d\pmb\sigma_1\,d\pmb\phi_2\,d\pmb\sigma_2\, e^{\int d1\,[L(\pmb\phi_1(1),\pmb\sigma_1(1))+L(\pmb\phi_2(1),\pmb\sigma_2(1))]} \end{equation} The same steps as in the derivation of the Euler characteristic follow. The result is the same as \eqref{eq:post.hubbard-strat}, but with the substitutions \begin{align} \mathbb Q\mapsto\begin{bmatrix} \mathbb Q_{11} & \mathbb Q_{12} \\ \mathbb Q_{21} & \mathbb Q_{22} \end{bmatrix} && \mathbb M\mapsto\begin{bmatrix} \mathbb M_1 \\ \mathbb M_2 \end{bmatrix} \end{align} where we have defined \begin{align} \mathbb Q_{ij}(1,2)=\frac1N\pmb\phi_i(1)\cdot\pmb\phi_j(2) && \mathbb M_i(1)=\frac1N\pmb\phi_i(1)\cdot\mathbf x_0 \end{align} Expanding the superindices and applying the Dirac $\delta$-functions implied by the Lagrange multipliers associated with the spherical constraint, we arrive at an expression \begin{equation} \frac1N\log\overline{\chi(\Omega)^2} =-\hat m_1-\hat m_2 -\frac\alpha2\log\frac{\det A_1}{\det A_2} -\frac{\alpha V_0^2}2\begin{bmatrix} 0 & 1 & 0 & 1 \end{bmatrix} A_1^{-1} \begin{bmatrix} 0 \\ 1 \\ 0 \\ 1 \end{bmatrix} +\frac12\log\frac{\det A_3}{\det A_4} \end{equation} with the matrices $A_1$, $A_2$, $A_3$, and $A_4$ defined by \begin{align} A_1&=\begin{bmatrix} D_{11}f'(1) & iR_{11}f'(1) & D_{12}f'(C_{12})+\Delta_{12}f''(C_{12}) & i R_{21}f'(C_{12}) \\ i R_{11}f'(1) & f(1) & i R_{12}f'(C_{12}) & f(C_{12}) \\ D_{12}f'(C_{12}) + \Delta_{12}f''(C_{12}) & iR_{12}f'(C_{12}) & D_{22} & iR_{22}f'(1) \\ iR_{21}f'(C_{12}) & f(C_{12}) & iR_{22}f'(1) & f(1) \end{bmatrix} \\ A_2&=\begin{bmatrix} 0 & R_{11}f'(1) & 0 & -G_{21}f'(C_{12}) \\ -R_{11}f'(1) & 0 & G_{12}f'(C_{12}) & 0 \\ 0 & -G_{12}f'(C_{12}) & 0 & R_{22}f'(1) \\ G_{21}f'(C_{12}) & 0 & -R_{22}f'(1) & 0 \end{bmatrix} \\ A_3&=\begin{bmatrix} 1-m_1^2 & i(R_{11}-m_1\hat m_1) & C_{12}-m_1m_2 & i(R_{21}-m_1\hat m_2) \\ i(R_{11}-m_1\hat m_1) & D_{11}+\hat m_1^2 & i(R_{12}-m_2\hat m_1) & D_{12}+\hat m_1\hat m_2 \\ C_{12}-m_1m_2 & i(R_{12}-m_2\hat m_1) & 1-m_2^2 & i(R_{22}-m_2\hat m_2) \\ i(R_{21}-m_1\hat m_2) & D_{12}+\hat m_1\hat m_2 & i(R_{22}-m_2\hat m_2) & D_{22}+\hat m_2^2 \end{bmatrix} \\ A_4&=\begin{bmatrix} 0 & R_{11} & 0 & -G_{21} \\ -R_{11} & 0 & G_{12} & 0 \\ 0 & -G_{12} & 0 & R_{22} \\ G_{21} & 0 & -R_{22} & 0 \end{bmatrix} \end{align} and where $\Delta_{12}=G_{12}G_{21}-R_{12}R_{21}$. The expression must be extremized over all the order parameters. We look for solutions in two regimes that are commensurate with the solutions found for the Euler characteristic. These correspond to $m_1=m_2=0$ and $C_{12}=0$, and $m_1=m_2=m^*$ and $C_{12}=1$. \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. 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_0 +\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} which is completely general for the spherical spin glasses with $M=1$. 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 supersymmetric 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 vanish. 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\hat\omega_0\,d\omega_1\,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_1r_d}C+I\right) \Bigg\} \end{aligned} \end{equation} 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 valid whether minima dominate or not. 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, 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 $\chi_0$ is \begin{equation} \chi_0(q)=\frac1{\hat\beta}[f''(q)^{-1/2}-\tilde r_d] \end{equation} the function implied by extremizing \eqref{eq:cont.action} over $\chi$. 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 $q_0<1$. 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 vanishes. Starting from this point, we take small steps in $s$ and $\lambda_s$, again 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. \bibliography{topology} \end{document}