diff options
author | Jaron Kent-Dobias <jaron@kent-dobias.com> | 2024-08-26 17:24:53 +0200 |
---|---|---|
committer | Jaron Kent-Dobias <jaron@kent-dobias.com> | 2024-08-26 17:24:53 +0200 |
commit | cf4dec33e62829f40240f29b484ea2f4c7c401e7 (patch) | |
tree | 7382fb31b6f94c6c189203b9f0c0a04ec047fe7e | |
parent | e10bf44c5261d7f025ffa9e88e0bef4b2783e7d3 (diff) | |
download | SciPostPhys_18_158-cf4dec33e62829f40240f29b484ea2f4c7c401e7.tar.gz SciPostPhys_18_158-cf4dec33e62829f40240f29b484ea2f4c7c401e7.tar.bz2 SciPostPhys_18_158-cf4dec33e62829f40240f29b484ea2f4c7c401e7.zip |
Lots of writing and figure work.
-rw-r--r-- | figs/action_1.pdf | bin | 20655 -> 20905 bytes | |||
-rw-r--r-- | figs/action_3.pdf | bin | 16245 -> 16245 bytes | |||
-rw-r--r-- | figs/dynamics_2.pdf | bin | 19291 -> 17774 bytes | |||
-rw-r--r-- | figs/dynamics_3.pdf | bin | 22214 -> 21696 bytes | |||
-rw-r--r-- | figs/phases_1.pdf | bin | 11318 -> 11318 bytes | |||
-rw-r--r-- | figs/phases_2.pdf | bin | 12012 -> 12012 bytes | |||
-rw-r--r-- | figs/phases_3.pdf | bin | 13050 -> 13049 bytes | |||
-rw-r--r-- | topology.bib | 14 | ||||
-rw-r--r-- | topology.tex | 605 |
9 files changed, 368 insertions, 251 deletions
diff --git a/figs/action_1.pdf b/figs/action_1.pdf Binary files differindex e4519a4..2a87561 100644 --- a/figs/action_1.pdf +++ b/figs/action_1.pdf diff --git a/figs/action_3.pdf b/figs/action_3.pdf Binary files differindex 749088d..54825b1 100644 --- a/figs/action_3.pdf +++ b/figs/action_3.pdf diff --git a/figs/dynamics_2.pdf b/figs/dynamics_2.pdf Binary files differindex 41ffece..a0fbcea 100644 --- a/figs/dynamics_2.pdf +++ b/figs/dynamics_2.pdf diff --git a/figs/dynamics_3.pdf b/figs/dynamics_3.pdf Binary files differindex b1763e4..b562de4 100644 --- a/figs/dynamics_3.pdf +++ b/figs/dynamics_3.pdf diff --git a/figs/phases_1.pdf b/figs/phases_1.pdf Binary files differindex 8f6c482..4a8c9c9 100644 --- a/figs/phases_1.pdf +++ b/figs/phases_1.pdf diff --git a/figs/phases_2.pdf b/figs/phases_2.pdf Binary files differindex bf16e76..a73666e 100644 --- a/figs/phases_2.pdf +++ b/figs/phases_2.pdf diff --git a/figs/phases_3.pdf b/figs/phases_3.pdf Binary files differindex 506ca73..e7a55fb 100644 --- a/figs/phases_3.pdf +++ b/figs/phases_3.pdf diff --git a/topology.bib b/topology.bib index 316df0f..8c17a2c 100644 --- a/topology.bib +++ b/topology.bib @@ -264,6 +264,20 @@ eprinttype = {arxiv} } +@article{Kent-Dobias_2023_How, + author = {Kent-Dobias, Jaron and Kurchan, Jorge}, + title = {How to count in hierarchical landscapes: a full solution to mean-field complexity}, + journal = {Physical Review E}, + publisher = {American Physical Society (APS)}, + year = {2023}, + month = {6}, + number = {6}, + volume = {107}, + pages = {064111}, + url = {https://doi.org/10.1103/PhysRevE.107.064111}, + doi = {10.1103/PhysRevE.107.064111} +} + @article{Kent-Dobias_2023_When, author = {Kent-Dobias, Jaron}, title = {When is the average number of saddle points typical?}, diff --git a/topology.tex b/topology.tex index 102dc6b..a7b68dc 100644 --- a/topology.tex +++ b/topology.tex @@ -34,7 +34,7 @@ \begin{abstract} We consider the set of solutions to $M$ random polynomial equations with - independent Gaussian coefficients on the $(N-1)$-sphere. When solutions + independent Gaussian coefficients and a target value $V_0$ on the $(N-1)$-sphere. When solutions exist, they form a manifold. We compute the average Euler characteristic of this manifold in the limit of large $N$, and find different behavior depending on the ratio $\alpha=M/N$. When $\alpha<\alpha_\text{onset}$, the @@ -110,7 +110,7 @@ tissues \cite{Fyodorov_2019_A, Fyodorov_2020_Counting, Montanari_2024_On, Kent-Dobias_2024_Conditioning, Kent-Dobias_2024_Algorithm-independent}. In each of these cases, the authors studied properties of the cost function \begin{equation} \label{eq:cost} - \mathscr C(\mathbf x)=\frac12\sum_{k=1}^MV_k(\mathbf x)^2 + \mathscr C(\mathbf x)=\frac12\sum_{k=1}^M\big[V_k(\mathbf x)-V_0\big]^2 \end{equation} which achieves zero only for configurations that satisfy all the constraints. Here we dispense with the cost function and study the set of solutions @@ -136,9 +136,188 @@ direct sum of the Betti numbers of $\Omega$. We find that for the varied cases we study, these two always coincide at the largest exponential order in $N$, putting strong constraints on the resulting topology and geometry. -\section{Results} +\section{Methods} -\subsection{Topology of solutions to many equations and the satisfiability transition} +\subsection{The average Euler characteristic} + +The Euler characteristic $\chi$ of a manifold is a topological invariant \cite{Hatcher_2002_Algebraic}. It is +perhaps most familiar in the context of connected compact orientable surfaces, where it +characterizes the number of handles in the surface: $\chi=2(1-\#)$ for $\#$ +handles. For general $d$, the Euler characteristic of the $d$-sphere is $2$ if $d$ is even and 0 if $d$ is odd. The canonical method for computing the Euler characteristic is done by +defining a complex on the manifold in question, essentially a +higher-dimensional generalization of a polygonal tiling. Then $\chi$ is given +by an alternating sum over the number of cells of increasing dimension, which +for 2-manifolds corresponds to the number of vertices, minus the number of +edges, plus the number of faces. + +Morse theory offers another way to compute the Euler characteristic using the +statistics of stationary points of a function $H:\Omega\to\mathbb R$ \cite{Audin_2014_Morse}. For +functions $H$ without any symmetries with respect to the manifold, the surfaces +of gradient flow between adjacent stationary points form a complex. The +alternating sum over cells to compute $\chi$ becomes an alternating sum over +the count of stationary points of $H$ with increasing index, or +\begin{equation} + \chi=\sum_{i=0}^N(-1)^i\mathcal N_H(\text{index}=i) +\end{equation} +Conveniently, we can express this abstract sum as an integral over the manifold +using a small variation on the Kac--Rice formula for counting stationary +points. Since the sign of the determinant of the Hessian matrix of $H$ at a +stationary point is equal to its index, if we count stationary points including +the sign of the determinant, we arrive at the Euler characteristic, or +\begin{equation} \label{eq:kac-rice} + \chi(\Omega)=\int_\Omega d\mathbf x\,\delta\big(\nabla H(\mathbf x)\big)\det\operatorname{Hess}H(\mathbf x) +\end{equation} +When the Kac--Rice formula is used to \emph{count} stationary points, the sign +of the determinant is a nuisance that one must take pains to preserve +\cite{Fyodorov_2004_Complexity}. Here we are correct to exclude it. + +We need to choose a function $H$ for our calculation. Because $\chi$ is +a topological invariant, any choice will work so long as it does not share some +symmetry with the underlying manifold, i.e., that it $H$ satisfies the Smale condition. Because our manifold of random +constraints has no symmetries, we can take a simple height function $H(\mathbf +x)=\mathbf x_0\cdot\mathbf x$ for some $\mathbf x_0\in\mathbb R^N$ with +$\|\mathbf x_0\|^2=N$. $H$ is a height function because when $\mathbf x_0$ is +used as the polar axis, $H$ gives the height on the sphere. + +We treat the integral over the implicitly defined manifold $\Omega$ using the +method of Lagrange multipliers. We introduce one multiplier $\omega_0$ to +enforce the spherical constraint and $M$ multipliers $\omega_k$ to enforce the vanishing of +each of the $V_k$, resulting in the Lagrangian +\begin{equation} \label{eq:lagrangian} + L(\mathbf x,\pmb\omega) + =H(\mathbf x)+\frac12\omega_0\big(\|\mathbf x\|^2-N\big) + +\sum_{k=1}^M\omega_k\big(V_k(\mathbf x)-V_0\big) +\end{equation} +The integral over $\Omega$ in \eqref{eq:kac-rice} then becomes +\begin{equation} \label{eq:kac-rice.lagrange} + \chi(\Omega)=\int_{\mathbb R^N} d\mathbf x\int_{\mathbb R^{M+1}}d\pmb\omega + \,\delta\big(\partial L(\mathbf x,\pmb\omega)\big) + \det\partial\partial L(\mathbf x,\pmb\omega) +\end{equation} +where $\partial=[\frac\partial{\partial\mathbf x},\frac\partial{\partial\pmb\omega}]$ +is the vector of partial derivatives with respect to all $N+M+1$ variables. +This integral is now in a form where standard techniques from mean-field theory +can be applied to calculate it. + +In order for certain Gaussian integrals in the following calculation to be +well-defined, it is necessary to treat instead the Lagrangian problem above +with $\pmb\omega\mapsto i\pmb\omega$. This transformation does not effect the +Dirac $\delta$ functions of the gradient, but it does change the determinant by +a factor of $i^{N+M+1}$. We will see that the result of the rest of the +calculation neglecting this factor is real. Since the Euler characteristic is +also necessarily real, this indicates an inconsistency with this transformation +when $N+M+1$ is odd. In fact, the Euler characteristic is always zero for +odd-dimensional manifolds. This is the signature of it in this problem. + +\subsubsection{Calculation of the average Euler characteristic} + +To evaluate the average of $\chi$ over the constraints, we first translate the $\delta$ functions and determinant to integral form, with +\begin{align} + \delta\big(\partial L(\mathbf x,\pmb\omega)\big) + &=\int\frac{d\hat{\mathbf x}}{(2\pi)^N}\frac{d\hat{\pmb\omega}}{(2\pi)^{M+1}} + e^{i[\hat{\mathbf x},\hat{\pmb\omega}]\cdot\partial L(\mathbf x,\pmb\omega)} + \\ + \det\partial\partial L(\mathbf x,\pmb\omega) + &=\int d\bar{\pmb\eta}\,d\pmb\eta\,d\bar{\pmb\gamma}\,d\pmb\gamma\, + e^{-[\bar{\pmb\eta},\bar{\pmb\gamma}]^T\partial\partial L(\mathbf x,\pmb\omega)[\pmb\eta,\pmb\gamma]} +\end{align} +where $\hat{\mathbf x}$ and $\hat{\pmb\omega}$ are ordinary vectors and +$\bar{\pmb\eta}$, $\pmb\eta$, $\bar{\pmb\gamma}$, and $\pmb\gamma$ are +Grassmann vectors. With these expressions substituted into +\eqref{eq:kac-rice.lagrange}, the result is a integral over an exponential +whose argument is linear in the random functions $V_k$. These functions can +therefore be averaged over, and the resulting expression treated with standard +methods. Details of this calculation can be found in Appendix~\ref{sec:euler}. +The result is the reduction of the average Euler characteristic to an expression of the +form +\begin{equation} + \overline{\chi(\Omega)} + =\int dR\,dD\,dm\,d\hat m\,g(R,D,m,\hat m)\,e^{N\mathcal S_\Omega(R,D,m,\hat m)} +\end{equation} +where $g$ is a prefactor subexponential in $N$, and $\mathcal S_\Omega$ is an effective action defined by +\begin{equation} + \begin{aligned} + \mathcal S_\Omega(R,D,m,\hat m\mid\alpha,V_0) + &=\hat m-\frac\alpha2\left[ + \log\left(1+\frac{f(1)D}{f'(1)R^2}\right) + +\frac{V_0^2}{f(1)}\left(1+\frac{f'(1)R^2}{f(1)D}\right)^{-1} + \right] \\ + &\hspace{7em}+\frac12\log\left( + 1+\frac{(1-m^2)D+\hat m^2-2Rm\hat m}{R^2} + \right) + \end{aligned} +\end{equation} +The remaining order parameters defined by the scalar products +\begin{align} + R=-i\frac1N\mathbf x\cdot\hat{\mathbf x} + && + D=\frac1N\hat{\mathbf x}\cdot\hat{\mathbf x} + && + m=\frac1N\mathbf x\cdot\mathbf x_0 + && + \hat m=-i\frac1N\hat {\mathbf x}\cdot\mathbf x_0 +\end{align} + +This integral can be evaluated by a saddle point method. For reasons we will +see, it is best to extremize with respect to $R$, $D$, and $\hat m$, leaving a +new effective action of $m$ alone. This can be solved to give +\begin{equation} + D=-\frac{m+R^*(m)}{1-m^2} \qquad \hat m=0 +\end{equation} +\begin{equation} + \begin{aligned} + R^*(m) + &=\frac{-m(1-m^2)}{2[f(1)-(1-m^2)f'(1)]^2} + \Bigg[ + \alpha V_0^2f'(1) + +(2-\alpha)f(1)\left(\frac{f(1)}{1-m^2}-f'(1)\right) \\ + &\quad+\operatorname{sgn}(m)\alpha\sqrt{ + \tfrac{4V_0^2}\alpha f(1)f'(1)\left[\tfrac{f(1)}{1-m^2}-f'(1)\right] + +\left[\tfrac{f(1)^2}{1-m^2}-\big(V_0^2+f(1)\big)f'(1)\right]^2 + } + \Bigg] + \end{aligned} +\end{equation} +\begin{equation} + \mathcal S_\Omega(m) + =-\frac\alpha2\bigg[ + \log\left( + 1-\frac{f(1)}{f'(1)}\frac{1+m/R^*}{1-m^2} + \right) + +\frac{V_0^2}{f(1)}\left( + 1-\frac{f'(1)}{f(1)}\frac{1-m^2}{1+m/R^*} + \right)^{-1} + \bigg] + +\frac12\log\left(-\frac m{R^*}\right) +\end{equation} +To finish evaluating the integral, this expression should be maximized with +respect to $m$. The order parameter $m$ is both physical and interpretable, as +it gives the overlap of the configuration $\mathbf x$ with the height axis +$\mathbf x_0$. Therefore, the value $m^*$ that maximizes this action can be +understood as the latitude on the sphere where most of the contribution to the +Euler characteristic is made. + +The action $\mathcal S_\Omega$ is extremized with respect to $m$ at $m^*=0$ or $m^*=-R^*$. +In the latter case, $m^*$ takes the value +\begin{equation} + m^*=\pm\sqrt{1-\frac{\alpha}{f'(1)}\big(V_0^2+f(1)\big)} +\end{equation} +and $\mathcal S_\Omega(m^*)=0$. However, when +\begin{equation} + V_0^2>V_\text{on}^2\equiv\frac{1-\alpha+\sqrt{1-\alpha}}\alpha f(1) +\end{equation} +$R^*(m^*)$ becomes complex and the solution is no longer valid. Likewise, when +\begin{equation} + V_0^2<V_\text{sh}^2\equiv\frac{2(1+\sqrt{1-\alpha})-\alpha}{\alpha}f(1)\left(1-\frac{f(1)}{f'(1)}\right) +\end{equation} +the maximum at $m=0$ becomes complex and that solution is invalid. Examples of +$\mathcal S_\Omega$ in difference regimes are shown in Fig.~\ref{fig:action}. +These transition values of the target $V_0$ correspond with transition values in $\alpha$ of +\begin{align} + \alpha_\text{on}=1-\left(\frac{V_0^2}{V_0^2+f(1)}\right)^2 + && + \alpha_\text{sh}=4V_0^2f(1)f'(1)\frac{f'(1)-f(1)}{\big((V_0^2+f(1))f'(1)-f(1)^2\big)^2} +\end{align} \begin{figure} \includegraphics{figs/action_1.pdf} @@ -146,13 +325,13 @@ putting strong constraints on the resulting topology and geometry. \includegraphics{figs/action_3.pdf} \caption{ - The effective action governing the as a function of the overlap + 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 $V_0$. In both + 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{Left:} With nonlinear + 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 @@ -160,9 +339,133 @@ putting strong constraints on the resulting topology and geometry. $V_0=V_\text{sh}\simeq1.394$, $m^*$ 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} + + + +\subsubsection{How to interpret the average Euler characteristic} + +It is not straightforward to directly use the average Euler characteristic to +infer something about the number of connected components in the set of +solutions. To understand why, a simple example is helpful. Consider the set of +solutions on the sphere $\|\mathbf x\|^2=N$ that satisfy the single quadratic +constraint +\begin{equation} + 0=\sum_{i=1}^N\sigma_ix_i^2 +\end{equation} +where each $\sigma_i$ is taken to be $\pm1$ with equal probability. If we take $\mathbf x$ to be ordered such that all terms with $\sigma_i=+1$ come first, this gives +\begin{equation} + 0=\sum_{i=1}^{N_+}x_i^2-\sum_{i=N_++1}^Nx_i^2 +\end{equation} +where $N_+$ is the number of terms with $\sigma_i=+1$. The topology of the resulting manifold can be found by adding and subtracting this constraint from the spherical one, which gives +\begin{align} + \frac12=\sum_{i=1}^{N_+}x_i^2 + \qquad + \frac12=\sum_{i=N_++1}^{N}x_i^2 +\end{align} +These are two independent equations for spheres of radius $1/\sqrt2$, one of +dimension $N_+$ and the other of dimension $N-N_+$. Therefore, the topology of +the configuration space is that of $S^{N_+-1}\times S^{N-N_+-1}$. The Euler +characteristic of a product space is the product of the Euler characteristics, +and so we have $\chi(\Omega)=\chi(S^{N_+-1})\chi(S^{N-N_+-1})$. + +What is the average value of the Euler characteristic over values of +$\sigma_i$? First, recall that the Euler characteristic of a sphere $S^d$ is 2 +when $d$ is even and 0 when $d$ is odd. When $N$ is odd, any value +of $N_+$ will result in one of the two spheres in the product to be +odd-dimensional, and therefore $\chi(\Omega)=0$, as is always true for +odd-dimensional manifolds. When $N$ is even, there are two possibilities: when $N_+$ is even then both spheres are odd-dimensional, while when $N_+$ is odd then both spheres are even-dimensional. +The number of terms $N_+$ with $\sigma_i=+1$ is distributed with the binomial distribution +\begin{equation} + P(N_+)=\frac1{2^N}\binom{N}{N_+} +\end{equation} +Therefore, the average Euler characteristic for even $N$ is +\begin{equation} + \overline{\chi(\Omega)} + =\sum_{N_+=0}^NP(N_+)\chi(S^{N_+-1})\chi(S^{N-N_+-1}) + =\frac4{2^N}\sum_{n=0}^{N/2}\binom{N}{2n} + =2 +\end{equation} +Thus we find the average Euler characteristic in this simple example is 2 +despite the fact that the possible manifolds resulting from the constraints +have characteristics of either 0 or 4. + +The many positive values the Euler characteristic can take is one problem, but +a larger one is the many negative values it can take. A manifold with many +handles will have a very negative Euler characteristic, and our calculation +relying on the saddle point of an exponential integral would be invalid. In the +regime of $\mathcal S_\Omega(m)$ where the action is complex-valued, is this +breakdown the result of a negative average characteristic, or is it the result +of another reason? It is difficult to answer this with the previous calculation +alone. However, in the next subsection we will make a complementary calculation +that rules out the negative Euler characteristic picture. Instead, we will see +that in the complex region, the large-deviation function in $m$ that $\mathcal +S_\Omega(m)$ represents breaks down due to the vanishingly small probability of +finding any stationary points. + +\subsection{Complexity of a linear test function} + +One way to definitely narrow possible interpretations of the average Euler +characteristic is to compute a complementary average. The Euler characteristic +is the alternating sum of numbers of critical points of different index. If +instead we make the direct sum +\begin{equation} + \mathcal N_H(\Omega)=\sum_{i=0}^N\mathcal N_H(\text{index}=i) + =\int_\Omega d\mathbf x\,\delta\big(\nabla H(\mathbf x)\big) + \,\big|\det\operatorname{Hess}H(\mathbf x)\big| +\end{equation} +we find the total number of stationary points. The formula is exactly the same +as that for the average Euler characteristic except for an absolute value sign +around the determinant of the Hessian. + +Understanding the number of stationary points as a function of latitude $m$ +will clarify the meaning of our effective action for the average Euler +characteristic. This is because the average number of stationary points is a +nonnegative number. If the region of complex $\mathcal S_\Omega$ has a +well-defined number of stationary points, it indicates that we are looking at a +situation with a negative average Euler characteristic. On the other hand, if +the average number of stationary points yields a complex value at some latitude +$m$, it must be because it is either to large or small in $N$ to be captured by +the calculation, e.g., that it behaves like $e^{N^2\Sigma}$ or +$e^{-N^2\Sigma}$. The following calculation indicates this second situation: +the region of complex action is due to a lack of stationary points to +contribute to the Euler characteristic at those latitudes. + +To compute the complexity, we follow a similar procedure to the Euler +characteristic. The main difference lies in how we treat the absolute value +function around the determinant. Following \cite{Fyodorov_2004_Complexity}, we +make use of the identity +\begin{equation} + \begin{aligned} + |\det A| + &=\lim_{\epsilon\to0}\frac{(\det A)^2}{\sqrt{\det(A+i\epsilon I)}\sqrt{\det(A-i\epsilon I)}} + \\ + &=\frac1{(2\pi)^N}\int d\bar{\pmb\eta}_1\,d\pmb\eta_1\,d\bar{\pmb\eta}_2\,d\pmb\eta_2\,d\mathbf a\,d\mathbf b\, + e^{ + -\bar{\pmb\eta}_1^TA\pmb\eta_1-\bar{\pmb\eta}_2^TA\pmb\eta_2 + -\frac12\mathbf a^T(A+i\epsilon I)\mathbf a-\frac12\mathbf b^T(A-i\epsilon I)\mathbf b + } + \end{aligned} +\end{equation} +for an $N\times N$ matrix $A$. Here $\bar{\pmb\eta}_1$, $\pmb\eta_1$, +$\bar{\pmb\eta}_2$, and $\pmb\eta_2$ are Grassmann vectors and $\mathbf a$ and +$\mathbf b$ are regular vectors. This introduces many new order parameters into +the problem, but this is a difficulty of scale rather than principle. With this +identity substituted for the usual determinant one, the problem can be solved +much as before. The details of this solution are relegated to +Appendix~\ref{sec:complexity.details}. The result is that, to largest order in +$N$, the logarithm of the average number of stationary points is the same as +the logarithm of the average Euler characteristic. + +\section{Results} + +\subsection{Topology of solutions to many equations and the satisfiability transition} + +The results of the previous sections indicate the following picture for the +topology of the solution manifold. + \begin{figure} \includegraphics[width=0.245\textwidth]{figs/connected.pdf} \includegraphics[width=0.245\textwidth]{figs/coexist.pdf} @@ -208,6 +511,30 @@ putting strong constraints on the resulting topology and geometry. \subsection{Topology of level sets of the spherical spin glasses and the dynamic threshold} +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 scaling $V_0=\sqrt\alpha E$. With a little algebra this procedure yields +\begin{equation} + E_\text{onset}=\pm\sqrt{2f(1)} +\end{equation} +\begin{equation} + E_\text{shatter}=\pm\sqrt{4f(1)\left(1-\frac{f(1)}{f'(1)}\right)} +\end{equation} +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 $f(q)=\frac12q^p$, +$E_\text{shatter}=\sqrt{2(p-1)/p}$, precisely the threshold energy in these +models. 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 generally satisfies $|E_\text{shatter}|\leq|E_\text{th}|$. + \begin{figure} \includegraphics{figs/dynamics_2.pdf} \hspace{-0.5em} @@ -226,87 +553,22 @@ putting strong constraints on the resulting topology and geometry. } \end{figure} -\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. +\section{Conclusions} -Morse theory offers another way to compute the Euler characteristic using the -statistics of stationary points of a function $H:\Omega\to\mathbb R$ \cite{Audin_2014_Morse}. For -functions $H$ without any symmetries with respect to the manifold, the surfaces -of gradient flow between adjacent stationary points form a complex. The -alternating sum over cells to compute $\chi$ becomes an alternating sum over -the count of stationary points of $H$ with increasing index, or -\begin{equation} - \chi=\sum_{i=0}^N(-1)^i\mathcal N_H(\text{index}=i) -\end{equation} -Conveniently, we can express this abstract sum as an integral over the manifold -using a small variation on the Kac--Rice formula for counting stationary -points. Since the sign of the determinant of the Hessian matrix of $H$ at a -stationary point is equal to its index, if we count stationary points including -the sign of the determinant, we arrive at the Euler characteristic, or -\begin{equation} \label{eq:kac-rice} - \chi(\Omega)=\int_\Omega d\mathbf x\,\delta\big(\nabla H(\mathbf x)\big)\det\operatorname{Hess}H(\mathbf x) -\end{equation} -When the Kac--Rice formula is used to \emph{count} stationary points, the sign -of the determinant is a nuisance that one must take pains to preserve -\cite{Fyodorov_2004_Complexity}. Here we are correct to exclude it. -We need to choose a function $H$ for our calculation. Because $\chi$ is -a topological invariant, any choice will work so long as it does not share some -symmetry with the underlying manifold, i.e., that it $H$ satisfies the Smale condition. Because our manifold of random -constraints has no symmetries, we can take a simple height function $H(\mathbf -x)=\mathbf x_0\cdot\mathbf x$ for some $\mathbf x_0\in\mathbb R^N$ with -$\|\mathbf x_0\|^2=N$. $H$ is a height function because when $\mathbf x_0$ is -used as the polar axis, $H$ gives the height on the sphere. +\cite{Franz_2016_The, Franz_2017_Universality, Franz_2019_Critical, Annesi_2023_Star-shaped, Baldassi_2023_Typical} -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} - 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. +\paragraph{Acknowledgements} +The authors thank Pierfrancesco Urbani for helpful conversations on these topics. -In order for certain Gaussian integrals in the following calculation to be -well-defined, it is necessary to treat instead the Lagrangian problem above -with $\pmb\omega\mapsto i\pmb\omega$. This transformation does not effect the -Dirac $\delta$ functions of the gradient, but it does change the determinant by -a factor of $i^{N+M+1}$. We will see that the result of the rest of the -calculation neglecting this factor is real. Since the Euler characteristic is -also necessarily real, this indicates an inconsistency with this transformation -when $N+M+1$ is odd. In fact, the Euler characteristic is always zero for -odd-dimensional manifolds. This is the signature of it in this problem. +\paragraph{Funding information} +JK-D is supported by a \textsc{DynSysMath} Specific Initiative of the INFN. -To evaluate the average of $\chi$ over the constraints, we first translate the $\delta$ functions and determinant to integral form, with -\begin{align} - \delta\big(\partial L(\mathbf x,\pmb\omega)\big) - &=\int\frac{d\hat{\mathbf x}}{(2\pi)^N}\frac{d\hat{\pmb\omega}}{(2\pi)^{M+1}} - e^{i[\hat{\mathbf x},\hat{\pmb\omega}]\cdot\partial L(\mathbf x,\pmb\omega)} - \\ - \det\partial\partial L(\mathbf x,\pmb\omega) - &=\int d\bar{\pmb\eta}\,d\pmb\eta\,d\bar{\pmb\gamma}\,d\pmb\gamma\, - e^{-[\bar{\pmb\eta},\bar{\pmb\gamma}]^T\partial\partial L(\mathbf x,\pmb\omega)[\pmb\eta,\pmb\gamma]} -\end{align} +\appendix + +\section{Details of the calculation of the average Euler characteristic} +\label{sec:euler} To make the calculation compact, we introduce superspace coordinates. Define the supervectors \begin{equation} @@ -317,8 +579,8 @@ To make the calculation compact, we introduce superspace coordinates. Define the The Euler characteristic can be expressed using these supervectors as \begin{equation} \begin{aligned} - \chi(\Omega) - &=\int d\pmb\phi\,d\pmb\sigma\,e^{\int d1\,L\big(\pmb\phi(1),\pmb\sigma(1)\big)} \\ + &\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) @@ -461,163 +723,17 @@ as setting everything depending on $\bar H$ and $H$ to zero. With these solution =\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{equation} -\begin{equation} - D=-\frac{m+R}{1-m^2} \qquad \hat m=0 -\end{equation} -\begin{equation} - \mathcal S(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) - +V_0^2\left(f(1)+\frac{f'(1)R^2}{D}\right)^{-1} - \right] - +\frac12\log\left( - 1+\frac{(1-m^2)D+\hat m^2-2Rm\hat m}{R^2} - \right) -\end{equation} -When $\alpha<\alpha_\text{onset}$ this potential has maxima at $\pm m^*$ with -$m^*=-R^*$ where its value is zero. -\begin{equation} - \alpha_\text{onset}=1-\left(\frac{V_0^2}{V_0^2+f(1)}\right)^2 -\end{equation} -\begin{equation} - \alpha_\text{shatter}=4V_0^2f(1)f'(1)\frac{f'(1)-f(1)}{\big((V_0^2+f(1))f'(1)-f(1)^2\big)^2} -\end{equation} - -\section{Complexity of the height function} - -\section{Implications for the spherical spin glasses} - -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 scaling $V_0=\sqrt\alpha E$. With a little algebra this procedure yields -\begin{equation} - E_\text{onset}=\pm\sqrt{2f(1)} -\end{equation} -\begin{equation} - E_\text{shatter}=\pm\sqrt{4f(1)\left(1-\frac{f(1)}{f'(1)}\right)} -\end{equation} -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 $f(q)=\frac12q^p$, -$E_\text{shatter}=\sqrt{2(p-1)/p}$, precisely the threshold energy in these -models. 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 generally satisfies $|E_\text{shatter}|\leq|E_\text{th}|$. - -\subsection{What does the average Euler characteristic tell us?} - -It is not straightforward to directly use the average Euler characteristic to -infer something about the number of connected components in the set of -solutions. To understand why, a simple example is helpful. Consider the set of -solutions on the sphere $\|\mathbf x\|^2=N$ that satisfy the single quadratic -constraint -\begin{equation} - 0=\sum_{i=1}^N\sigma_ix_i^2 -\end{equation} -where each $\sigma_i$ is taken to be $\pm1$ with equal probability. If we take $\mathbf x$ to be ordered such that all terms with $\sigma_i=+1$ come first, this gives -\begin{equation} - 0=\sum_{i=1}^{N_+}x_i^2-\sum_{i=N_++1}^Nx_i^2 -\end{equation} -where $N_+$ is the number of terms with $\sigma_i=+1$. The topology of the resulting manifold can be found by adding and subtracting this constraint from the spherical one, which gives -\begin{align} - \frac12=\sum_{i=1}^{N_+}x_i^2 - \qquad - \frac12=\sum_{i=N_++1}^{N}x_i^2 -\end{align} -These are two independent equations for spheres of radius $1/\sqrt2$, one of -dimension $N_+$ and the other of dimension $N-N_+$. Therefore, the topology of -the configuration space is that of $S^{N_+-1}\times S^{N-N_+-1}$. The Euler -characteristic of a product space is the product of the Euler characteristics, -and so we have $\chi(\Omega)=\chi(S^{N_+-1})\chi(S^{N-N_+-1})$. - -What is the average value of the Euler characteristic over values of -$\sigma_i$? First, recall that the Euler characteristic of a sphere $S^d$ is 2 -when $d$ is even and 0 when $d$ is odd. When $N$ is odd, any value -of $N_+$ will result in one of the two spheres in the product to be -odd-dimensional, and therefore $\chi(\Omega)=0$, as is always true for -odd-dimensional manifolds. When $N$ is even, there are two possibilities: when $N_+$ is even then both spheres are odd-dimensional, while when $N_+$ is odd then both spheres are even-dimensional. -The number of terms $N_+$ with $\sigma_i=+1$ is distributed with the binomial distribution -\begin{equation} - P(N_+)=\frac1{2^N}\binom{N}{N_+} -\end{equation} -Therefore, the average Euler characteristic for even $N$ is -\begin{equation} - \overline{\chi(\Omega)} - =\sum_{N_+=0}^NP(N_+)\chi(S^{N_+-1})\chi(S^{N-N_+-1}) - =\frac4{2^N}\sum_{n=0}^{N/2}\binom{N}{2n} - =2 -\end{equation} -Thus we find the average Euler characteristic in this simple example is 2 -despite the fact that the possible manifolds resulting from the constraints -have characteristics of either 0 or 4. - +\section{Calculation of the prefactor of the average Euler characteristic} +\label{sec:prefactor} -\cite{Franz_2016_The, Franz_2017_Universality, Franz_2019_Critical, Annesi_2023_Star-shaped, Baldassi_2023_Typical} - -\section{Average number of stationary points of a test function} - -\subsection{Behavior with extensively many constraints} - -\subsection{Behavior with finitely many constraints} - -\begin{equation} - Mv_0^2=\frac4{f''(1)}\left[ - f'(1)-f(1)-2\frac{f(1)}{f'(1)}f''(1) - \right]\left[ - \frac{f(1)}{f'(1)}f''(1)-2\big(f'(1)-f(1)\big) - \right] -\end{equation} - -\begin{equation} - Mv_0^2=\frac{f'(1)^2}{f''(1)} -\end{equation} +\section{Details of the calculation of the average number of stationary points} +\label{sec:complexity.details} -\section{Interpretation of our results} +\section{The quenched shattering energy in {\oldstylenums 1}\textsc{frsb} models} +\label{sec:1frsb} -\[ - E=\frac1{\hat\omega_1}\frac12\left[ - \hat\omega_1^2f(C)+(2\omega_1\hat\omega_1R-\omega_1^2D)f'(C)+\omega_1^2(R^2-G^2)f''(C)+\log\det\frac{CD+R^2}{G^2} - \right] -\] -$D=\beta R$, $R=r_dI$, $G=-r_dI$ -\[ - E=\frac1{\hat\omega_1}\frac12\left[ - \hat\omega_1^2f(C)+(2\omega_1\hat\omega_1-\omega_1^2\beta)r_df'(1)+\log\det(\beta r_d^{-1}C+I) - \right] -\] -$\beta=\hat\omega_1/\omega_1$ -\[ - E=\frac1{\hat\omega_1}\frac12\left[ - \hat\omega_1^2f(C)+\omega_1\hat\omega_1r_df'(1)+\log\det(\hat\omega_1\omega_1^{-1}r_d^{-1}C+I) - \right] -\] -$z=r_d\omega_1$ -\[ - E=\frac1{\hat\omega_1}\frac12\left[ - \hat\omega_1^2f(C)+\hat \omega_1 zf'(1)+\log\det(\hat\omega_1C/z+I) - \right] -\] -\[ - \log\chi - =-\hat\omega_1 E - +\frac12[2\omega_1\hat\omega_1r_d-\omega_1^2d_d]f'(1) - +\frac12\int_0^1dq\,\left[ - \hat\omega_1^2f''(q)\chi(q) - +\frac1{\chi(q)+r_d^2/d_d} - \right] - -\frac12\log r_d^2 -\] -\[ - 0=\frac12\hat\omega_1^2f''(q)-\frac12\frac1{[\chi(q)+r_d^2/d_d]^2} -\] +\cite{Kent-DObias_2023_How} \[ \chi_0(q)=\frac1{\hat\omega_1}f''(q)^{-1/2}-\frac{r_d^2}{d_d} \] @@ -651,19 +767,6 @@ $z=r_d\omega_1$ d_d=-\frac{1+r_d}{\int dq\,\chi(q)}r_d \] -\paragraph{Acknowledgements} -The authors thank Pierfrancesco Urbani for helpful conversations on these topics. - -\paragraph{Funding information} -JK-D is supported by a \textsc{DynSysMath} Specific Initiative of the INFN. - -\appendix - -\section{Calculation of the prefactor of the average Euler characteristic} -\label{sec:prefactor} - -\section{The quenched shattering energy in {\oldstylenums 1}\textsc{frsb} models} -\label{sec:1frsb} \bibliography{topology} |