Zum Inhalt
Erschienen in:

Open Access 2025 | OriginalPaper | Buchkapitel

Reachability for Nonsmooth Systems with Lexicographic Jacobians

verfasst von : Chenxi Ji, Huan Zhang, Sayan Mitra

Erschienen in: Tools and Algorithms for the Construction and Analysis of Systems

Verlag: Springer Nature Switzerland

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Das Kapitel vertieft die Komplexität der Erreichbarkeitsanalyse für nichtglatte dynamische Systeme, die durch kontinuierliche, aber nicht differenzierbare Funktionen gekennzeichnet sind. Traditionelle Methoden haben mit diesen Systemen zu kämpfen, weil es an nicht differenzierbaren Punkten keine klar definierten Jakobiner gibt, was zu konservativen Übernäherungen führt. Die Autoren schlagen einen bahnbrechenden Algorithmus vor, der lexikographische Differenzierung zur Berechnung direktionaler Derivate nutzt und wertvolle Informationen erster Ordnung selbst an nicht differenzierbaren Punkten liefert. Dieser Ansatz umgeht die Notwendigkeit einer expliziten Berechnung erreichbarer Grenzwerte, eine bedeutende Barriere in der hybriden Erreichbarkeitsanalyse. Der Algorithmus wird durch experimentelle Ergebnisse an 12 nicht glatten dynamischen Benchmark-Modellen validiert, die eine bis zu 4-mal höhere Genauigkeit aufweisen als hochmoderne Werkzeuge wie CORA und JuliaReach. Das Kapitel untersucht auch die Anwendung dieser Methode auf neuronale ODEs der ReLU, einer besonders herausfordernden Klasse nichtglatter Systeme, und demonstriert ihr Potenzial, die Erreichbarkeitsanalyse komplexer dynamischer Systeme voranzutreiben.

1 Introduction

Systems described by ordinary differential equations (ODEs), \(\frac{dx}{dt} = f(x)\), where the function f is continuous but not differentiable are called nonsmooth dynamical systems. Nonsmooth systems arise in various real-world scenarios. For instance, in describing the motion of autonomous vehicles f could encompass the physics of the car as well as the decision logic of the autonomy software. A toy example of this type is illustrated at the end of this section. Sharp decisions, whether implemented through if-then-else conditions or ReLU activation functions, introduce nonsmoothness. This paper is concerned with reachability analysis for such nonsmooth dynamical systems.
Reachability analysis aims to compute the set of states that can be reached from a given set of initial states through trajectories of a dynamical system. This reachable set can be used to automatically verify safety with respect to uncertainties in the initial conditions. Various algorithms and software tools have been developed for the reachability analysis of linear, nonlinear, and hybrid dynamical systems (see, for example,  [12, 31, 32, 35]). Extending these ideas to nonsmooth systems presents two barriers.
First, with the hybrid systems [25] point of view1, a nonsmooth system can be partitioned into several smooth subsystems (or modes) glued together by mode transitions (see Figure 1). Reachability analysis of such hybrid models proceeds by performing reachability within each smooth subdomain, followed by carefully propagating the reachable set across the mode transitions or the decision boundaries [4, 7, 11, 39]. The solutions from different initial states may encounter the mode boundaries (e.g., \(x_f-x_r=d\pm \frac{a}{c}\) in Example 1) at different points or at different times. Consequently, the entire set of reachable boundary states from the initial set must be computed before propagating it forward into the next mode. In practice, this reachable boundary set is determined by computing reachable states in small time increments, checking nonempty intersections with the boundary conditions, and then aggregating these non-empty intersections. Some variation of this approach underlies most of the hybrid verification tools such as CORA [13], JuliaReach [7, 9, 10], Flow* [11], C2E2 [16], SpaceEx [23], DryVR [21], and Verse [32]. This process often results in overtly conservative over-approximations due to the “wrapping effect” of combining pieces of the reachable boundary set [5]. Our proposed method effectively addresses this challenge by avoiding explicit computation of the reachable boundary sets.
Secondly, the standard method for reachability analysis of smooth nonlinear ODEs uses the sensitivity of solutions to initial conditions [13, 17]. Given a time-varying (sensitivity) function \(\beta :\mathbb {R}_{\ge 0}\times \mathbb {R}_{\ge 0} \rightarrow \mathbb {R}\) that bounds the distance between two solutions as a function of their initial distance (i.e. \(|x_1(t) - x_2(t)|_p \le \beta (|x_1(0) - x_2(0)|_p,t)\)), the reachable set can be approximated by a sphere centered around a reference solution, with a radius defined by \(\beta \) [18]. For smooth ODEs, the sensitivity function \(\beta \) can be computed from the Jacobian \(J_f\) of system dynamics or the associated matrix measures [20, 34]. This strategy is not applicable for nonsmooth functions, because the Jacobian matrix \(J_f\) is undefined at the non-differentiable points.
To tackle the above challenges, we introduce a novel algorithm that leverages directional derivatives, which provide valuable first-order information even at non-differentiable points. Our approach employs the notion of lexicographic differentiation which was introduced by Nesterov in [36]. This method computes a sequence of directional derivatives defined by a matrix \(M\) (see Definition 2), transforming the original nonsmooth function into a differentiable one and extracting first-order information in specific directions from the newly constructed differentiable function. The Lexicographic Jacobian \(J_L\) offers directional first-order information that can be computed using a generalized chain rule [30], unlike the Clarke Jacobian [14]—which is an alternative set-valued Jacobian used for nonsmooth systems.
The algorithm provides sound over-approximations of the reachable set by propagating the initial set using an optimized exponential sensitivity function, calculated over a set of lexicographic Jacobians. Lexicographic Jacobians allow us to circumvent the problem of explicitly computing the reachable boundary set as is needed with the hybrid approach. The correctness proof (Theorem 1) extends the fundamental theorem of calculus to nonsmooth functions (see Lemma 1), which may be of independent interest. A key innovation of our algorithm is the use of polytopic approximations for sets of L-Jacobians, replacing the traditional rectangular approximations [20]. This significantly improves the precision of reachability analysis and allows the linear approximations to leverage efficient linear bound propagation tools like CROWN [41, 45, 46].
A particularly challenging class of nonsmooth systems are ReLU Neural ODEs—systems with dynamics defined by neural networks (NN) with ReLU activations. These models have found applications in modeling irregular time series data, system identification, and generative tasks [37, 38, 44]. Because of the nonsmoothness of these ODEs the Taylor model-based approaches for Neural ODEs with smooth activations (e.g., tanh or sigmoid) cannot be directly applied here [26, 27, 33, 43]. Secondly, the large number of mode boundaries make the computation of reachable boundary set and the hybrid approach challenging [3, 8]. It is important to note here that the neural ODE reachability requires the bounds to be propagated through an integral of the NN-defined function, which is different from the NN Control System (NNCS) reachability [19, 22, 28, 42, 47], which only requires bounds to be propagated through a NN at discrete times. In this paper, we bound the L-Jacobian of any ReLU network by taking the product of the bounds on the Clarke Jacobians of its individual layers. This approach is validated by the generalized chain rule for nonsmooth functions [30] and the fact that the L-Jacobian of any ReLU layer is included within its Clarke Jacobian (Proposition 3). We subsequently utilize this bound in our framework to obtain the reachability of ReLU neural ODEs.
We present a software implementation that uses CROWN to compute linear bounds on sets of Lexicographic Jacobians and then solves an optimal expansion rate satisfying a quadratic constraint over each element within the bounds. The initial set is then expanded by an exponential sensitivity function based on this rate and evolution time, yielding a sound over-approximation of the reachable set. Our experimental results on 12 benchmark nonsmooth dynamical models show that this algorithm outperforms state-of-the-art reachability tools (CORA [1] and JuliaReach [7]) on most problems. On average, our method is 50% more accurate, and in some cases, up to 4 times more accurate. Our approach enables reachability analysis of Neural ODEs with ReLU activations for models up to 12 dimensions and 200 neurons-an essential class of nonsmooth systems. To our knowledge, this is the first method capable of such analysis.
An Illustrative Example. Consider a car following a leader car moving at a constant speed. Let \(x_f\), \(x_r\), \(v_f\), and \(v_r\) be the positions and velocities of the front and rear cars. The dynamics of this system with a simple rule to maintain safe following distance can be written as \(\dot{x}_f = v_f, \dot{v}_f = 0, \dot{x}_r = v_r\), and
where \(a>0\) is the maximum braking acceleration, \(d>0\) is a threshold distance, and \(c>0\) is a constant coefficient. The above figure illustrates the vector \((\dot{v}_f, \dot{v}_r)\) in the \(x_f\)-\(x_r\) space. The vector subspace outside the region bounded by the orange lines represents cases where \(\dot{v}_r\) is a constant value (\(\pm a\)), while the subspace within this region represents the case where \(\dot{v}_r\) is a linear function.
Figure 1 shows the reachable sets for the above car following example with \(d=7\) and an initial set centered at \((x_f,x_r)=(1,-10)\) in the \(x_f-x_r\) domain computed by our method, CORA [3], and JuliaReach [7]. For similar initial sets, our method verifies safety over a time horizon of 13 seconds (staying below red line, where \(x_f > x_r\)). In contrast, CORA and JuliaReach, fail to verify safety even for a 6-second time horizon and required significantly more time for computation (beyond 60 minutes). This comparison underscores the advantages of our method over hybrid methods for nonsmooth systems.
Fig. 1.
Reachable sets in the \(x_f\)-\(x_r\) domain computed by our method, CORA, and JuliaReach. Red Line: Safety Boundary; Orange Line: Nonsmooth Hyperplanes; Black Ellipsoids/Zonotopes: Initial Sets; Blue Ellipsoids/Zonotopes: Reachable Sets. Black lines in left subfigure: Sampled trajectories, indicating our competed sets are valid over-approximation of exact reachable sets.

2 Preliminaries

Notations. For \(x\in \mathbb {R}^n\) and a positive definite matrix P, define \(||x||_P=\sqrt{x^TPx}\). The set \(E_{P,c}(x)=\{x' \in \mathbb {R}^n \ | \ ||x'-x||_P^2\le c\}\) denotes the ellipsoid centered at x with shape matrix P and size c. The standard 2-norm is \(||x||_2=||x||_I\) and \(B_c(x) = E_{I,c}(x)\) represents the Euclidean ball centered at x of radius c. We also define \(B_c(S)=\bigcup _{x\in S} B_c(x)\), for a set \(S \subseteq \mathbb {R}^n\). For a set \(S\subseteq \mathbb {R}^n\), \({dia}(S) =\sup _{x,x' \in S}(||x-x'||_2)\). For a matrix A, \(||A||_2\) denotes its 2-norm; \(A\succ 0\) and \(A\succeq 0\) indicate that A is a positive definite and positive semi-definite, respectively. For a differentiable function \(f:\mathbb {R}^n \rightarrow \mathbb {R}^n\), \(Jf \in \mathbb {R}^{n \times n}\) is the Jacobian of f. Given a set of \(n\times n\) matrices \(\mathcal {A}\), its convex hull is defined as, \(conv(\mathcal {A})=\{sA_1+(1-s)A_2|A_1,A_2\in \mathcal {A}, s\in [0,1]\}\), which collects matrices representable as linear combinations of those in \(\mathcal {A}\).
Dynamical System. The continuous evolution of a system is mathematically described as a dynamical system. Consider an n-dimensional dynamical system:
$$\begin{aligned} \dot{x}=f(x) \end{aligned}$$
(1)
where \(f:\mathbb {R}^n \rightarrow \mathbb {R}^n\) is a locally Lipschitz continuous function describing the continuous evolution of the state variables. A nonsmooth system implies that f is continuous and piecewise continuously differentiable (\(PC^1\)), but not necessarily differentiable. A solution or trajectory of the system starting at \(x_0\) is a function \(x:\mathbb {R}_{\ge 0} \rightarrow \mathbb {R}^n\) where \(x(0)=x_0\) and for any \(t\in \mathbb {R}_{\ge 0}\), x(t) satisfies Equation (1).
Reachability. For the dynamical system defined as Eq. (1), a state \( y \in \mathbb {R}^n \) is called reachable within time interval \([t_1, t_2]\) from initial set \(\mathcal {X}\) if there exists \( x_0 \in \mathcal {X} \) and a time \( t \in [t_1, t_2] \) such that \( y = x(t) \) with \( x(0)=x_0 \). The set of all such states is the reach set, denoted \( \text {Reach}(\mathcal {X}, [t_1, t_2]) \). Similarly, \( \text {Reach}(\mathcal {X}, t_1) \) represents the states reachable precisely at \( t_1 \). Reachability analysis seeks to over-approximate \( \text {Reach}(\mathcal {X}_0, [0, T]) \) for an initial set \( \mathcal {X}_0 \subset \mathbb {R}^n \) over \([0, T]\). While traditional methods use Jacobian-based sensitivity functions for smooth systems, for nonsmooth systems the Jacobians are undefined at the non-differentiable points, and those methods are unsound. The Clarke Jacobian [14] is a related notion that is well-defined for nonsmooth functions.
Clarke Jacobian. For a locally Lipschitz continuous function \( f: \mathbb {R}^n \rightarrow \mathbb {R}^m \), the Clarke Jacobian, denoted \( J_C f(x) \), is the convex hull of Jacobians near \( x \) where \( f \) is differentiable:
$$ J_C f(x) = \text {conv} \left\{ \lim _{x_k \rightarrow x} J f(x_k) \mid f \text { is differentiable at } x_k \right\} . $$
It is important to note that Clarke Jacobian does not satisfy the chain rule, complicating its computation for ReLU networks, where layer functions are explicit but the overall expression is not. We introduce the L-Jacobian next, which follows the generalized chain rule [30] and allows for layer-by-layer computation.

3 Lexicographical Derivatives

Lexicographical differentiation was introduced by Nesterov in [36] and the rules for automatic differentiation were later developed by Barton et al. in [6]. In this section, we fix \(f:\mathbb {R}^n \rightarrow \mathbb {R}^m\) to be a continuous function.
Definition 1
The directional derivative of f at \(x\in \mathbb {R}^n\) in the direction \(d\in \mathbb {R}^n\) is given by the following limit, provided it exists:
$$\begin{aligned} f'(x;d)=\lim \limits _{\alpha \rightarrow 0^+}\frac{f(x+\alpha d)-f(x)}{\alpha }. \end{aligned}$$
For example, the ReLU function \(f=\max (x,0)\) is not differentiable at 0, but its directional derivative \(f'(0;d) = 0\) for \(d \le 0\) and \(f'(0;d) = 1\) for \(d > 0\) exist. A function is directionally differentiable at x if \(f'(x;d)\) exists for all directions d. The notion of lexicographic smoothness (L-smoothness) uses a sequence of higher-order directional derivatives, at points given by the rows of a matrix M.
Definition 2
If the following sequence of directional derivatives are well-defined for any \(k\in \mathbb {N}\) and \(M=[m_1,m_2,\dots ,m_k]\in \mathbb {R}^{n\times k}\) then f is lexicographically smooth (L-smooth) at \(x\in \mathbb {R}^n\).
$$\begin{aligned} & f_{x,M}^{(0)}:\mathbb {R}^n\rightarrow \mathbb {R}^n:d\rightarrow f'(x;d),\\ & f_{x,M}^{(1)}:\mathbb {R}^n\rightarrow \mathbb {R}^n:d\rightarrow [f_{x,M}^{(0)}]'(m_1;d),\\ & \vdots \\ & f_{x,M}^{(k)}:\mathbb {R}^n\rightarrow \mathbb {R}^n:d\rightarrow [f_{x,M}^{(k-1)}]'(m_k;d). \end{aligned}$$
Here, \(f'(x;d)\) represents the standard directional derivative of f at state x in the direction d, while \(f_{x,M}^{(0)}\) refers to the mapping that assigns each input d to its corresponding value of \(f'(x;d)\). Similarly, \([f_{x,M}^{(i)}]'(m_{i+1};d)\) presents the standard directional derivative of \(f_{x,M}^{(i)}\) estimated at \(m_{i+1}\) in the direction d, while \(f_{x,M}^{(i+1)}\) refers to the mapping that assigns each input d to its corresponding value of \([f_{x,M}^{(i)}]'(m_{i+1};d)\).
Consider an example function \(f(x_1,x_2)=|x_1^2-x_2^2|\), which is non-differentiable at \(x_0=(1,1)^T\). For different \(2 \times 2\) M matrices, the directional derivatives of f as a function of the direction \(d=(d_1,d_2)\) are shown in Table 1 alongside the L-and Clarke Jacobians.
Table 1.
Directional derivatives, L and Clarke Jacobians of \(f(x_1,x_2)=|x_1^2-x_2^2|\).
https://static-content.springer.com/image/chp%3A10.1007%2F978-3-031-90653-4_18/MediaObjects/652619_1_En_18_Tab1_HTML.png
Piece-wise continuously differentiable functions \(PC^1\) and their compositions and integrals, are L-smooth [30]. Furthermore, the solution to any L-smooth ODE is also L-smooth [29, 30, 36]. The next proposition is one of the key properties of lexicographic differentiation and it states that for any k and a full row rank \(n\times k\) matrix M, the \(k^{ t h}\) directional derivative \(f_{x,M}^{(k)}(d)\) is linear w.r.t. d [36]. For the above example, indeed \(f_{x,M}^{(2)}(d)\) is linear, but only for the first three M’s which are full row rank.
Proposition 1
For an L-smooth f and a full row rank matrix \(M\in \mathbb {R}^{n\times k}\), the \(k^{ th }\) directional derivative \(f^{(k)}_{x,M}(d)\) is linear w.r.t direction d.
For any linear function, and particularly \(f^{(k)}_{x,M}\) in Proposition 1, its regular derivative is well-defined, which then justifies the following definition of Lexicographic derivative or the L-Jacobian for L-smooth functions.
Definition 3
Given a L-smooth function f and a full row rank matrix \(M\in \mathbb {R}^{n\times k}\), the L-Jacobian of f at x in the direction given by M is
$$\begin{aligned} J_Lf(x;M)=J f^{(k)}_{x;M}(\vec {0})\in \mathbb {R}^{m\times n}. \end{aligned}$$
(2)
That is, the L-Jacobain of f corresponds to normal derivative of \(k^{ th }\) directional function \(f^{(k)}_{x;M}\). Since \(f^{(k)}_{x;M}\) is linear w.r.t its input d according to Proposition 1, its normal Jacobian function therefore must be a constant function. Thus, we replace the arbitrary input d with a particular direction, namely \(\vec {0}\). We introduce the definition of the Lexicographic subdifferential, denoted as \(\partial _L f(\mathcal {X})\), which captures the generalized derivative information of the function f over a set \(X \subseteq \mathbb {R}^n\) in terms of the L-Jacobian [30].
Definition 4
For an L-smooth function f, a set \(\mathcal {X}\subseteq \mathbb {R}^n\), its Lexicographic subdifferential (L-subdifferential) over \(\mathcal {X}\) is a set of \(m\times n\) matrices given by:
$$\begin{aligned} \partial _L f(\mathcal {X}) = \bigcup _{x \in \mathcal {X}} \left\{ \begin{array}{ll} \{J_L f(x;M) \mid M \in \mathbb {R}^{n \times n}, \det (M) \ne 0\} & \text {if } f \text {not differentiable at } x, \\ J f(x) & \text {otherwise}. \end{array} \right. \end{aligned}$$
This subdifferential captures the generalized derivative information across both smooth and nonsmooth regions of X. The following proposition from [30] relates the standard directional derivative with L-Jacobian.
Proposition 2
For any L-smooth f and full row rank matrix \(M\in \mathbb {R}^{n\times k}\) with first column d, the directional derivative of f at x in the direction d is \(f'(x;d)=J_Lf(x;M)\cdot d\).
For an L-smooth function f, we derived the following lemma, which is an analog of the fundamental theorem of calculus in terms of the L-Jacobian \(J_Lf(x;M)\).
Lemma 1
For any L-smooth function f, for any full row rank matrix \(M\in \mathbb {R}^{n\times k}\) with first column d, for any \(x \in \mathbb {R}^n\),
$$\begin{aligned} f(x+d)-f(x)=\big (\int _0^1 J_Lf(x+sd;M)ds\big )\cdot d. \end{aligned}$$
(3)
Proof
We use i’s in the subscript to denote the \(i^{th}\) component of a vector or a function. For any \(s\in [0,1]\), \(i\in \{1,\cdots ,m\}\), we define \(g_i(s)=f_i(x+sd)\). Then we have \(\forall s\in [0,1]\),
$$\begin{aligned} f_i(x+d)-f_i(x)=g_i(1)-g_i(0),\quad f'_i(x+sd;d)=g'_i(s;1). \end{aligned}$$
(4)
Since f is L-smooth, \(f_i\) is Lipschitz over closed interval \([x,x+d]\). Thus, \(g_i\) is also Lipschitz over \(s\in [0,1]\). All Lipschitz continuous functions over a closed interval have bounded variation. It follows that \(g_i\) is a bounded variation function over \(s\in [0,1]\). By Corollary 6 in Chapter 6.3 of [40], since \(g_i\) is of bounded variation over [0, 1], \(g_i\) is differentiable almost everywhere w.r.t s over open interval (0, 1) and is integrable over [0, 1]. Thus, we have \(g_i(1)-g_i(0)=\int _0^1 g'_i(s) ds=\int _0^1 g'_i(s;1)ds\). Considering all \(i\in \{1,\cdots ,m\}\), we have \(g(1)-g(0)=\int _0^1 g'(s;1)ds\). By substituting g back to f, we have \(f(x+d)-f(x)=\int _0^1 f'(x+sd;d)ds\). By applying Proposition 2 (\(f'(x+sd;d)=J_Lf(x+sd;M)\cdot d\)), the result follows.    \(\square \)
We close this section with a proposition relating the Clarke Jacobian with the lexicographic subdifferential. This relationship is relevant for the reachability analysis of ReLU neural ODEs (see Section 5).
Proposition 3
The L-subdifferential of an \( n \)-dimensional ReLU function \(\sigma :\mathbb {R}^n \rightarrow \mathbb {R}^n\) satisfies \(\partial _L \sigma (\mathcal {X}) \subseteq J_C(\mathcal {X})\) for any \(\mathcal {X} \subseteq \mathbb {R}^n\).
Proposition 3 indicates that the Clarke Jacobian over-approximates the L-subdifferential of any \( n \)-dimensional ReLU layer. According to the generalized chain rule for L-smooth functions [30], the L-Jacobian of a ReLU network is the product of the L-Jacobians of its individual layers, enabling a layer-by-layer bound on the network’s L-subdifferential. By applying Proposition 3, tools that bound the Clarke Jacobian of ReLU layers can conservatively over-approximate the network’s L-subdifferential. It is important to note that this layer-by-layer bound does not guarantee an over-approximation of the network’s Clarke Jacobian due to the inapplicability of chain rule for the Clarke Jacobian.

4 Reachability from Lexicographic Subdifferentials

In the rest of this paper, we fix \(f:\mathbb {R}^n\rightarrow \mathbb {R}^n\) to be a L-smooth (Definition 2) function. In this section, our goal is to compute the reachable set of the dynamical system (1) from \(\mathcal {X}_0\) over time \(t\in [0,T]\). Our method, first computes a numerical simulation \(x(\cdot )\) from a specific initial state \(x(0) \in \mathcal {X}_0\), and then expands a set around \(x(\cdot )\) containing all trajectories from \(\mathcal {X}_0\). In Section 4.1, we introduce a method to bound distance between any two solutions (trajectories) by computing bounds on quadratic constraints of L-subdifferentials. In Sections 4.2 and 4.3, we present two methods to relax number of quadratic constraints for cases where the L-subdifferential can be linearly approximated. In Section 4.4, we present the overall reachability algorithm.

4.1 Bounding the Displacement between Solutions

Proposition 4
Given any convex compact set \(\mathcal {X}^*\subseteq \mathbb {R}^n\), if \(\partial _L f(\mathcal {X}^*)\) is a convex set, then \(\forall x_1,x_2\in \mathcal {X}^*\), full row rank \(M\in \mathbb {R}^{n\times k}\), then
$$\begin{aligned} \int _0^1 J_Lf(x_1+s(x_2-x_1);M)ds \in \partial _L f(\mathcal {X}). \end{aligned}$$
(5)
Proof
Since \(\mathcal {X}^*\) is a convex set, for any \(x_1,x_2\in \mathcal {X}^*\), any \(s\in [0,1]\), \(J_Lf(x_1+s(x_2-x_1);M)\in \partial _L f(\mathcal {X})\). Further, since \(\partial _L f(\mathcal {X})\) is convex, the result follows.    \(\square \)
In the following, we denote \(y(\cdot )=x_1(\cdot )-x_2(\cdot )\) as the difference between two solutions of  (1). Theorem 1 states that the norm of y(t) can be bounded in term of its initial value y(0) and time t.
Theorem 1
Given \(\mathcal {X}_0\subseteq \mathbb {R}^n\), let \(\mathcal {X}^*\) be a compact convex set such that \(\forall x(0)\in \mathcal {X}_0\), \(\forall t\in [0,T]\), \(x(t)\in \mathcal {X}^*\). Assume \(\partial _L(\mathcal {X}^*)\)is convex, and there exists a positive definite matrix \(P\in \mathbb {R}^{n \times n}\) and a real \(\gamma \) such that \(A^TP+PA\preceq \gamma P\) for all \(A\in \partial _L(\mathcal {X}^*)\). Then, \(\forall x_1(0),x_2(0)\in \mathcal {X}_0\), \(\forall t\in [0,T]\),
$$\begin{aligned} ||x_1(t)-x_2(t)||_P\le e^{\frac{\gamma }{2}t}||x_1(0)-x_2(0)||_P. \end{aligned}$$
(6)
Proof
Let us fix \(x_1(0),x_2(0)\in \mathcal {X}_0\) and denote \(y(t)=x_1(t)-x_2(t)\). For a positive definite matrix \(P\succ 0\), \(||y(t)||_P^2=y^T(t)Py(t)\). By differentiating \(||y(t)||_P^2\), we have that for any \(t\in [0,T]\),
$$\begin{aligned} \frac{d||y(t)||_P^2}{dt} = & \frac{d(y^T(t)Py(t))}{dt} = \dot{y}^T(t)Py(t)+y^T(t)P\dot{y}(t), \end{aligned}$$
(7)
According to Lemma 1, \(\dot{y}(t)\) can be written as:
$$\begin{aligned} \dot{y}(t)=\left( \int _0^1 J_Lf\big ( x_1(t)+sy(t));M \big ) ds \right) \cdot y(t), \end{aligned}$$
(8)
According to Proposition 4, the parenthetic term of the RHS of (8) belongs to \(\partial _L f(\mathcal {X}^*)\). By combining Equations (7), (8), we have
$$\begin{aligned} \frac{d||y(t)||_P^2}{dt} = y^T(t)(A^T(t)P+PA(t))y(t), \end{aligned}$$
(9)
for some \(A(t)\in \partial _L f(\mathcal {X}^*)\). Using the assumption about \(\gamma \), it follows that \( \frac{d||y(t)||_P^2}{dt} \le \gamma y^T(t)Py(t) = \gamma ||y(t)||_P^2\). By applying Grönwall’s inequality, we have \(\forall t\in [0,T]\), \(||y(t)||_P\le ||y(0)||_P e^{\frac{\gamma }{2}t}\). After substituting \(y(t)=x_1(t)-x_2(t)\), the result follows.    \(\square \)
Theorem 1 bounds the displacement of two solutions in terms of their initial distance, the exponential factor \(\gamma \), and the matrix P which defines the norm. Thus, we can compute an ellipsoid centered at \(x_1(t)\) (computed by numerical integration of  (1)), to get an over-approximation of the reachable states at time t. The conservativeness of this over-approximation depends on the choice of the matrix P and \(\gamma \). Ideally, we would like to pick the optimal \(P \succ 0\) to minimize \(\gamma \), which can be written as:
$$\begin{aligned} \begin{aligned} & \min _{\gamma \in \mathbb {R},P \succ 0} \gamma \ \ \mathsf {subject \ to}\ \ A^TP+PA\preceq \gamma P, \quad \forall A\in \partial _L f(\mathcal {X}^*).\\ \end{aligned} \end{aligned}$$
(10)

4.2 Relaxation to Finite Number of Constraints

The optimization problem in (10) contains over uncountably infinite constraints—one for each A in \(\partial _L f(\mathcal {X}^*)\). Next, we present a method that replaces this infinite collection of constraints with a relaxed finite collection.
Lemma 2
Given a set of matrices \(\mathcal {A}\subset \mathbb {R}^{n\times n}\), assume there exists a set of finite matrices \(\mathcal {V}\subset \mathbb {R}^{n\times n}\) such that \(\mathcal {A}\subseteq conv(\mathcal {V})\). If there exist \( \gamma \in \mathbb {R}\) and positive definite \(P\in \mathbb {R}^{n\times n}\) such that \(V^TP+PV\preceq \gamma P\) for all \(V\in \mathcal {V}\), then
$$\begin{aligned} A^TP+PA\preceq \gamma P,\quad \forall A\in \mathcal {A}. \end{aligned}$$
(11)
Lemma 2 suggests that the optimization problem (10) can be simplified as the following optimization problem.
$$\begin{aligned} \begin{aligned} & \min _{\gamma \in \mathbb {R},P\succ 0} \gamma \ \ \mathsf {subject \ to} \ \ V^TP+PV\preceq \gamma P, \quad \forall V\in \mathcal {V}. \end{aligned} \end{aligned}$$
(12)
Note that if \(\gamma \) is fixed, the optimization problem (12) becomes a semi-definite programming (SDP), and a feasible solution can be obtained by SDP solver. Thus, we can solve (12) by a linear search strategy, where a SDP is solved at each step. This strategy may not lead to optimal \(\gamma \), but we can get a sufficient close estimation of optimal \(\gamma \) by starting linear search from a negative enough lower bound \(\underline{\gamma }\) with a small enough search step \(\varDelta \gamma \).
We present a second method to solve optimization problem (10), which contains only one constraint and calculates a remainder term. This makes it much more computationally efficient, especially on higher dimensional examples.
Lemma 3
Given a set of matrices \(\mathcal {A}\subset \mathbb {R}^{n\times n}\), assume there exists a set of finite matrices \(\mathcal {V}\subset \mathbb {R}^{n\times n}\) such that \(\mathcal {A}\subseteq conv(\mathcal {V})\). For any \(A_c\in \mathbb {R}^{n\times n}\), if there exists \(\hat{\gamma }\in \mathbb {R}\) and positive definite \(P\in \mathbb {R}^{n\times n}\) such that \(A_c^TP+PA_c\preceq \hat{\gamma }P\), then
$$\begin{aligned} A^TP+PA\preceq (\hat{\gamma }+ \frac{\delta }{\lambda _{min(P)}})P, \quad \forall A\in \mathcal {A}, \end{aligned}$$
(13)
where \(\delta =\max \limits _{V'\in \mathcal {V}'} ||V'||_2\), \(\mathcal {V}'=\{(V-A_c)^TP+P(V-A_c)|V\in \mathcal {V}\}\) and \(\lambda _{min}(P)\) refers the minimum eigenvalue of P.
Lemma 3 suggests that the optimization problem (10) can be simplified as the following optimization problem.
$$\begin{aligned} \begin{aligned} & \min _{\hat{\gamma }\in \mathbb {R},P\succ 0} \hat{\gamma } \ \ \mathsf {subject \ to} \ \ A_c^TP+PA_c\preceq \hat{\gamma } P \end{aligned} \end{aligned}$$
(14)
To minimize the expansion rate \(\hat{\gamma }+\frac{\delta }{\lambda _{min}(P)}\), we heuristically select \(A_c\) as the average value of \(\mathcal {V}\). Additionally, by leveraging Lemma 3 with Theorem 1, we derive a sensitivity function that provides a bound on the displacement between trajectories:
$$\begin{aligned} ||y(t)||_P\le e^{(\frac{\hat{\gamma }}{2}+\frac{\delta }{2\lambda _{min}(P)})t}||y(0)||_P. \end{aligned}$$
(15)
This center matrix based method is significantly less computationally resource-intensive than the vertices based method (Lemma 2) at the price of decreasing the accuracy, due to the positive error term \(\delta \) that is added to \(\hat{\gamma }\) in (13). In practice, we want to make the compact sets \(\mathcal {X}^*\) (in Theorem 1) small so that \(\delta \) and \(\hat{\gamma }+\frac{\delta }{\lambda _{min}(P)}\) (the expansion rate) remains small.

4.3 Linear Over-Approximations of L-Subdifferentials

When we apply either Lemmas 2 or 3 to simplify the semidefinite constraints in Theorem 1, \(\mathcal {A}\) refers to \(\partial _L f(\mathcal {X}^*)\). Notice that both lemmas require constructing a finite set \(\mathcal {V}\) whose convex hull contains \(\partial _L f(\mathcal {X}^*)\), which can be reframed as first constructing a polytopic superset of \(\partial _L f(\mathcal {X}^*)\) by linear approximation, then extracting its vertices for \(\mathcal {V}\).
Linear approximation bounds each output coordinate with two linear functions (lower and upper) concerning the input variables. This method offers better accuracy than constant approximations and superior computational efficiency compared to polynomial approximations, especially when dealing with complex computation graphs like those found in neural networks. In this work, we utilize CROWN, which takes the computation graph of the L-Jacobian and computes linear bounds on each nonsmooth or nonlinear component. It then propagates these bounds layer by layer to derive the linear approximation of \(\partial _L f(\mathcal {X}^*)\) over a specified input region.
Specifically, for an L-smooth function \( f: \mathbb {R}^n \rightarrow \mathbb {R}^n \) over a bounded set \(\mathcal {X}^* \subset \mathbb {R}^n\), CROWN generates matrices \(\underline{A}, \overline{A} \in \mathbb {R}^{n^2 \times n}\) and vectors \(\underline{b}, \overline{b} \in \mathbb {R}^{n^2}\) so that for all \(i\) and \(x \in \mathcal {X}^*\),
$$\begin{aligned} \underline{A}_{i}x + \underline{b}_i \le [\text {vec}(G)]_i \le \overline{A}_{i}x + \overline{b}_i, \quad \forall G \in \partial _L f(x). \end{aligned}$$
(16)
Here, \([\text {vec}(G)]_i\) denotes the \(i\)-th element of the vectorized matrix \(G\).
Next, we present an algorithm that constructs a polytopic superset of L-subdifferential of f and extracts its vertices using the linear approximations derived from CROWN over the region \(\mathcal {X}^*\).
Algorithm 1
Compute Over-approximation of L-subdifferentials (COL)
The input to Algorithm 1 is a convex compact polytope \(\mathcal {X}^*\subset \mathbb {R}^n\) and the output is a set of finite matrices \(\mathcal {V}\subset \mathbb {R}^{n\times n}\). Algorithm 1 proceeds as follows: At Line 1, it utilizes CROWN to compute coefficients of linear bounds \(\{\underline{A}, \overline{A},\underline{b}, \overline{b}\}\), which satisfies Inequality (16). At Line 2, it constructs a vector set, \(vec_V\), containing \(n^2\) length vectors derived from the linear equations \(V_i = A_ix + b_i\). Here, \(A\) can be either \(\underline{A}\) or \(\overline{A}\), \(b\) can be either \(\underline{b}\) or \(\overline{b}\), and x belong to the vertices of input polytope \(\mathcal {X}^*\). Each vector in \(vec_V\) corresponds to evaluations of the L-subdifferential at the vertices of the input polytope \(\mathcal {X}^*\). Finally, at Line 3, it reshapes \(vec_V\) into \(\mathcal {V}\) by reorganizing its elements into a set of \(n \times n\) matrices.
Proposition 5
If \(\mathcal {X}^*\subset \mathbb {R}^n\) is a convex compact polytope, Algorithm 1 outputs \(\mathcal {V}\subset \mathbb {R}^{n\times n}\) such that \(conv(\mathcal {V})\) is an over-approximation of \(\partial _L f(\mathcal {X}^*)\).

4.4 Reachability Algorithm

Theorem 1 outlines a method to compute sensitivity functions over given convex compact initial set within bounded time horizon. For unstable systems, computing a single sensitivity function over the entire time duration may lead to significant over-approximations due to the large set of L-subdifferentials. To mitigate this, we divide the time horizon into smaller intervals and compute piecewise sensitivity functions. In the following, we present an algorithm based on Theorem 1 and Lemma 2 to compute an over-approximation of \( {Reach}(\mathcal {X}_0, [0,T]) \) for system (1).
Algorithm 2
Reachable Set Computation
Algorithm 2 takes the inputs: (1) an initial set \(\mathcal {X}_0\subset \mathbb {R}^n\); (2) time bound \(T>0\); (3) Lipschitz constant \(L_f\) of f (can be replaced by a local Lipschitz constant for each time interval); (4) the number of time subintervals k. The output is a set R contains \(Reach(\mathcal {X}_0,[0,T])\).
Initially, the overall R is initialized to be an empty set; the time step for each iteration \(\varDelta t\) is \(\frac{T}{k}\). An ellipsoid \(E_{P_0,c_0}(x_0)\), centered at \(x_0\) with shape matrix \(P_0\) and size \(c_0\), is computed to ensure that it contains \(\mathcal {X}_0\), is computed (Line 2). In each iteration, the algorithm sets \(t_{i}\) as the endpoint of current time interval; and computes a solution of (1) from initial state \(x_0\) evaluated at time instance \(t_i\), denoted as \(x_i\); (Line 5). It then computes a convex compact set \(\mathcal {X}_i\) containing the solution starting from \(x_0\) and evolving within \([t_{i-1},t_i]\), and enlarging set \(\mathcal {X}_i\) by a factor of \(d_{i-1}e^{L_ft}\) to obtain \(\mathcal {X}^*_i\) (Line 6); and computes a set of finite matrices \(\mathcal {V}\) by Algorithm 1 (Line 7). Locally optimal \(\gamma _i\) and \(P_i\) is calculated by solving optimization problem (12) using linear search strategy (Line 8). To ensure that the ellipsoid in shape \(P_{i-1}\) is included within the ellipsoid in new shape \(P_i\), it computes the minimum temporary scalar \(c'_{i-1}\) and updates it to get \(c_i\) accordingly (Line 9). Subsequently, the diameter \(d_i\) of the ellipsoid at time \(t_i\) is computed (Line 10). \(d'\) is assigned as the maximum diameter of ellipsoid-shape over-approximations of reachable sets evolving within \([t_{i-1},t_i]\) (Line 11). Finally, the reachable set for time interval \([t_{i-1},t_i]\) is obtained by expanding \(\mathcal {X}_i\) with half of \(d'_i\) (Line 12), and the resulting set is added to the overall reachable set R. This process continues until all time intervals are processed. Theorem 2 confirms the soundness of Algorithm 2 and Proposition 6 indicates that the accuracy of computed reachable set can be improved by finer subdivisions of \(\mathcal {X}_0\).
Theorem 2
For any L-smooth function f, initial set \(X_0\) and time duration T, the output R of Algorithm 2 over-approximates \(Reach(\mathcal {X}_0,[0,T])\).
Proposition 6
For each i, as \(dia(\mathcal {X}_i)\rightarrow 0\) the bloating factor \(d_i\rightarrow 0\).
This proposition indicates that the over-approximation error from expansion can be minimized by reducing the initial set’s uncertainty. As the initial set \(\mathcal {X}_0\) approaches zero, the size \(c_0\) of the initial ellipsoid in shape matrix \(P_0\) and the bloating factors \(d_i\), both converge to zero.

5 Experimental Results

We implemented Algorithm 1 using CROWN to compute linear bounds on L-subdifferentials and Algorithm 2 with CVXPY [15] to solve Optimization 12. In this section, we discuss the results of reachability analysis performed using this implementation compared with several other tools on a set of benchmark problems.
Benchmarks. For nonsmooth systems, we use three types of benchmark models: (1) hybrid models that are also nonsmooth dynamical systems (e.g. the CarFollowing model); (2) nonsmooth systems created by adding switch conditions to standard smooth system benchmarks from the ARCH competition [24] (Switch-\(\langle .\rangle \) models); (3) small ReLU neural ODEs with manually-set weights (SmallNN-\(\langle .\rangle \) models). For ReLU neural ODEs, we use models with randomly initialized weights (randNN-\(\langle .\rangle \) models) and neural ODEs trained to approximate standard smooth benchmarks (Approx-\(\langle .\rangle \) models). For smooth systems, we focus on benchmark from the ARCH competition. Full descriptions are given in Appendix.
Baselines. We compared our method against CORA [1] and JuliaReach [10], two leading hybrid verification tools from the ARCH competition. We also compare our results against the original Jacobian-based method of [20] on smooth models. The results suggest that our method has significant advantages across the board.
Metrics. We evaluate performance of reachability tools on time efficiency (RT) and conservativeness of computed results (F/I and A/I). RT refers to running time, F/I denotes the ratios of the final reachable set’s volume over the initial set’s volume (\(\frac{vol( Reach (\mathcal {X}_0,T))}{vol(\mathcal {X}_0)}\)); A/I represents the ratio of the average volume of the all reachable sets over the initial set’s volume (\(\frac{\sum _{i=1}^k vol( Reach (\mathcal {X}_0,t_i))}{k\times vol(\mathcal {X}_0)}\)). These metrics are appropriate for comparison across tools with different reach set representations.
Fig. 2.
Reachable sets computed by our tools, CORA, and JuliaReach on three nonsmooth models: SmallNN-1, Switch-Robertson, and Switch-Laub-Loomis. Projections of initial sets (black) and reachable sets (blue) and nonsmooth hyperplanes (black lines) are shown. With similar initial sets, our method yields the least conservative reachable sets.
Table 2.
Comparison of reachability tools (Ours, CORA, and JuliaReach) on nonsmooth models. Our method outperforms CORA and JuliaReach on most cases in terms of accuracy (lower A/I and F/I), especially for higher-dimensional systems or those with increased mode transitions. Dim: System Dimension; ID: Initial Diameter; TH: Time Horizon; A/I: Average-to-Initial Volume Ratio; F/I: Final-to-Initial Volume Ratio; RT: Running Time.
https://static-content.springer.com/image/chp%3A10.1007%2F978-3-031-90653-4_18/MediaObjects/652619_1_En_18_Tab2_HTML.png
Tightest Over-Approximations on Nonsmooth Models Results are shown in Table 2, and Figure 2 visualizes the reachable sets computed by our algorithm (Algorithm 2), CORA, and JuliaReach for three models: SimpleNN-1, Switch-Robertson, and Switch-Vanderpol. Table 2 indicates that on lower-dimensional models (Dim \(\le 3\)), our algorithm generally outperforms CORA and JuliaReach in both accuracy and runtime, except for two models with brief transitions (Switch-Bruss and Switch-Lotka-Volterra, Line 1 and 3 in Table 2). For higher-dimensional models (Dim \(\ge 4\)), our algorithm consistently provides a more accurate over-approximation, achieving up to 4 times better results on the CarFollowing model (Line 9 in Table 2). This advantage is expected, as traditional hybrid reachability methods become more conservative on higher-dimensional systems due to over-approximation at transition boundaries. Overall, these results demonstrate that our approach provides greater accuracy and efficiency, particularly for high-dimensional systems and those with numerous transitions and nonsmooth boundaries.
First Reachability Method for ReLU Neural ODEs According to Proposition 3, we use CROWN [41] to compute the product of bounds on the Clarke Jacobians of individual layers, which serves as a conservative bound on the L-subdifferential of the entire ReLU neural ODE. We then integrate this bound into the framework outlined in Algorithms 1 and Algorithm 2 to achieve reachability.
As shown in Table 3 and Figure 3, the computed sets (blue ellipsoids) encompass all sampled trajectories (black), illustrating their role as over-approximations within acceptable time consumption. Additionally, their close proximity to the sampled trajectories demonstrates low conservativeness on low-dimensional models (Dim \(\le 3\)).
Fig. 3.
Reachable sets of ReLU Neural ODEs computed using our method. Projections of initial sets (black), reachable sets (blue) and sampled trajectories (black lines) are displayed. The computed reachsets contain system trajectories, reinforcing our conclusion that the computation results represent over-approximations.
Table 3.
Results of Reachability for ReLU Neural Models using our algorithm. This table demonstrates that our method scales effectively to ReLU neural systems of up to 12 dimensions with 200 neurons, which significantly surpasses typical hybrid systems in the number of transitions. NN Structure: Layered Architecture of ReLU Neural Models; Dim: System dimension; ID: Initial diameter; TH: Time horizon; WAM: Weight Assignment Method (trained for training; normal for normal distribution; uniform for uniform distribution); A/I: Average-to-Initial Volume Ratio; F/I: Final-to-Initial Volume Ratio; RT: Running time.
https://static-content.springer.com/image/chp%3A10.1007%2F978-3-031-90653-4_18/MediaObjects/652619_1_En_18_Tab3_HTML.png
Tighter Over-Approximation on Smooth Models. Our Algorithm 2, using CROWN, achieves tighter Jacobian bounds for smooth dynamical systems, resulting in less conservative reachable sets than the method from [20]. As shown in Table 4, our approach consistently demonstrates improved accuracy (smaller A/I and F/I) across all examples, especially for higher-dimensional models (Lines 7-9).
Table 4.
Comparison of reachability tools (our method and the method in [20]) on benchmark smooth systems. Results show that our approach achieves higher accuracy on all benchmarks. Dim: System Dimension; ID: Initial Diameter; TH: Time Horizon; A/I: Average-to-Initial Volume Ratio; F/I: Final-to-Initial Volume Ratio; RT: Running Time.
https://static-content.springer.com/image/chp%3A10.1007%2F978-3-031-90653-4_18/MediaObjects/652619_1_En_18_Tab4_HTML.png

6 Conclusion and Future Directions

We have presented a new algorithm for reachability analysis of a general class of nonsmooth dynamical systems using the recently developed notion of lexicographic differentiation. Our implementation employs CROWN for computing linear approximations of lexicographic Jacobians, which are used to bound the sensitivity of solutions to initial conditions. This method, which does not require hybridization or partitioning the state space into differentiable subdomains, introduces a new approach for analyzing a broad class of hybrid systems and neural ODEs. Our experiments suggest that this method can be more accurate than alternative hybrid reachability algorithms.
Applications of Lexicographic differentiation in formal verification and controller synthesis are worth exploring further. For instance, higher order directional derivatives could enable Taylor model-based reachability analysis for nonsmooth systems. More detailed, potentially compositional analysis of neural ODEs and dynamical systems with neural network controllers is another direction of emerging interest.

Acknowledgments

This work was supported by a research grant from the AFOSR MURI program (FA9550-23-1-0337) with a project entitled HyDDRA: Hybrid Dynamics - Deconstruction and Aggregation. Huan Zhang is supported in part by the AI2050 program at Schmidt Sciences (AI2050 Early Career Fellowship) and NSF (IIS-2331967).
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
Fußnoten
1
Every nonsmooth dynamical system can be represented as a hybrid system by partitioning its state space into differentiable subspaces with transitions occurring at nonsmooth boundaries. However, not all hybrid systems can be considered nonsmooth systems, as hybrid systems may include discontinuous vector fields, whereas in this paper, nonsmooth systems specifically refer to systems with continuous but non-differentiable vector fields.
 
Literatur
1.
Zurück zum Zitat M. Althoff. An introduction to cora 2015. In Proc. of the 1st and 2nd Workshop on Applied Verification for Continuous and Hybrid Systems, pages 120–151, December 2015. M. Althoff. An introduction to cora 2015. In Proc. of the 1st and 2nd Workshop on Applied Verification for Continuous and Hybrid Systems, pages 120–151, December 2015.
2.
Zurück zum Zitat M. Althoff and D. Grebenyuk. Implementation of interval arithmetic in cora 2016. In Proc. of the 3rd Workshop on Applied Verification for Continuous and Hybrid Systems, pages 91–105, February 2017. M. Althoff and D. Grebenyuk. Implementation of interval arithmetic in cora 2016. In Proc. of the 3rd Workshop on Applied Verification for Continuous and Hybrid Systems, pages 91–105, February 2017.
3.
Zurück zum Zitat M. Althoff, D. Grebenyuk, and N. Kochdumper. Implementation of taylor models in cora 2018. In Proc. of the 5th Workshop on Applied Verification for Continuous and Hybrid Systems, pages 145–173, September 2018. M. Althoff, D. Grebenyuk, and N. Kochdumper. Implementation of taylor models in cora 2018. In Proc. of the 5th Workshop on Applied Verification for Continuous and Hybrid Systems, pages 145–173, September 2018.
4.
Zurück zum Zitat R. Alur, C. Courcoubetis, T. A. Henzinger, and P.-H. Ho. Hybrid automata: An algorithmic approach to the specification and verification of hybrid systems. In Hybrid Systems, volume 736, pages 209–229. Springer, 1992. R. Alur, C. Courcoubetis, T. A. Henzinger, and P.-H. Ho. Hybrid automata: An algorithmic approach to the specification and verification of hybrid systems. In Hybrid Systems, volume 736, pages 209–229. Springer, 1992.
5.
Zurück zum Zitat E. Asarin, T. Dang, G. Frehse, A. Girard, C. Le Guernic, and O. Maler. Recent progress in continuous and hybrid reachability analysis. In 2006 IEEE Conference on Computer Aided Control System Design, 2006 IEEE International Conference on Control Applications, 2006 IEEE International Symposium on Intelligent Control, pages 1582–1587. IEEE, 2006. E. Asarin, T. Dang, G. Frehse, A. Girard, C. Le Guernic, and O. Maler. Recent progress in continuous and hybrid reachability analysis. In 2006 IEEE Conference on Computer Aided Control System Design, 2006 IEEE International Conference on Control Applications, 2006 IEEE International Symposium on Intelligent Control, pages 1582–1587. IEEE, 2006.
6.
Zurück zum Zitat P. I. Barton, K. A. Khan, P. Stechlinski, and H. A. Watson. Computationally relevant generalized derivatives: theory, evaluation and applications. Optimization Methods and Software, 33(4-6):1030–1072, 2018. P. I. Barton, K. A. Khan, P. Stechlinski, and H. A. Watson. Computationally relevant generalized derivatives: theory, evaluation and applications. Optimization Methods and Software, 33(4-6):1030–1072, 2018.
7.
Zurück zum Zitat S. Bogomolov, M. Forets, G. Frehse, K. Potomkin, and C. Schilling. Juliareach: a toolbox for set-based reachability. In Proceedings of the 22nd ACM International Conference on Hybrid Systems: Computation and Control, pages 39–44, 2019. S. Bogomolov, M. Forets, G. Frehse, K. Potomkin, and C. Schilling. Juliareach: a toolbox for set-based reachability. In Proceedings of the 22nd ACM International Conference on Hybrid Systems: Computation and Control, pages 39–44, 2019.
8.
Zurück zum Zitat S. Bogomolov, M. Forets, G. Frehse, K. Potomkin, and C. Schilling. Reachability analysis of linear hybrid systems via block decomposition. IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, 39(11):4018–4029, Nov. 2020. S. Bogomolov, M. Forets, G. Frehse, K. Potomkin, and C. Schilling. Reachability analysis of linear hybrid systems via block decomposition. IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, 39(11):4018–4029, Nov. 2020.
9.
Zurück zum Zitat S. Bogomolov, M. Forets, G. Frehse, F. Viry, A. Podelski, and C. Schilling. Reach set approximation through decomposition with low-dimensional sets and high-dimensional matrices. In Proceedings of the 21st International Conference on Hybrid Systems: Computation and Control (part of CPS Week), pages 41–50, 2018. S. Bogomolov, M. Forets, G. Frehse, F. Viry, A. Podelski, and C. Schilling. Reach set approximation through decomposition with low-dimensional sets and high-dimensional matrices. In Proceedings of the 21st International Conference on Hybrid Systems: Computation and Control (part of CPS Week), pages 41–50, 2018.
10.
Zurück zum Zitat S. Bogomolov, M. Forets, and K. Potomkin. Case study: Reachability and scalability in a unified combat-command-and-control model. In International Conference on Reachability Problems, pages 52–66. Springer, 2020. S. Bogomolov, M. Forets, and K. Potomkin. Case study: Reachability and scalability in a unified combat-command-and-control model. In International Conference on Reachability Problems, pages 52–66. Springer, 2020.
11.
Zurück zum Zitat X. Chen, E. Ábrahám, and S. Sankaranarayanan. Flow*: An analyzer for non-linear hybrid systems. In CAV, pages 258–263. Springer, 2013. X. Chen, E. Ábrahám, and S. Sankaranarayanan. Flow*: An analyzer for non-linear hybrid systems. In CAV, pages 258–263. Springer, 2013.
12.
Zurück zum Zitat X. Chen and S. Sankaranarayanan. Reachability analysis for cyber-physical systems: Are we there yet? In NASA Formal Methods Symposium, pages 109–130. Springer, 2022. X. Chen and S. Sankaranarayanan. Reachability analysis for cyber-physical systems: Are we there yet? In NASA Formal Methods Symposium, pages 109–130. Springer, 2022.
13.
Zurück zum Zitat A. Chutinan and B. H. Krogh. Verification of polyhedral-invariant hybrid automata using polygonal flow pipe approximations. LNCS, 1569:76–91, 1999. A. Chutinan and B. H. Krogh. Verification of polyhedral-invariant hybrid automata using polygonal flow pipe approximations. LNCS, 1569:76–91, 1999.
14.
Zurück zum Zitat F. H. Clarke. Generalized gradients and applications. Transactions of the American Mathematical Society, 205:247–262, 1975. F. H. Clarke. Generalized gradients and applications. Transactions of the American Mathematical Society, 205:247–262, 1975.
15.
Zurück zum Zitat S. Diamond and S. Boyd. Cvxpy: A python-embedded modeling language for convex optimization. The Journal of Machine Learning Research, 17(1):2909–2913, 2016. S. Diamond and S. Boyd. Cvxpy: A python-embedded modeling language for convex optimization. The Journal of Machine Learning Research, 17(1):2909–2913, 2016.
16.
Zurück zum Zitat P. Duggirala, S. Mitra, M. Viswanathan, and M. Potok. C2e2: A verification tool for stateflow models. In TACAS, pages 68–82. Springer, 2015. P. Duggirala, S. Mitra, M. Viswanathan, and M. Potok. C2e2: A verification tool for stateflow models. In TACAS, pages 68–82. Springer, 2015.
17.
Zurück zum Zitat P. S. Duggirala, S. Mitra, and M. Viswanathan. Verification of annotated models from executions. In EMSOFT, 2013. P. S. Duggirala, S. Mitra, and M. Viswanathan. Verification of annotated models from executions. In EMSOFT, 2013.
18.
Zurück zum Zitat P. S. Duggirala, S. Mitra, and M. Viswanathan. Verification of annotated models from executions. In EMSOFT, page 26. IEEE Press, 2013. P. S. Duggirala, S. Mitra, and M. Viswanathan. Verification of annotated models from executions. In EMSOFT, page 26. IEEE Press, 2013.
19.
Zurück zum Zitat S. Dutta, X. Chen, and S. Sankaranarayanan. Reachability analysis for neural feedback systems using regressive polynomial rule inference. In Proceedings of the 22nd ACM International Conference on Hybrid Systems: Computation and Control, HSCC ’19, pages 157–168, New York, NY, USA, 2019. ACM. S. Dutta, X. Chen, and S. Sankaranarayanan. Reachability analysis for neural feedback systems using regressive polynomial rule inference. In Proceedings of the 22nd ACM International Conference on Hybrid Systems: Computation and Control, HSCC ’19, pages 157–168, New York, NY, USA, 2019. ACM.
20.
Zurück zum Zitat C. Fan, J. Kapinski, X. Jin, and S. Mitra. Locally optimal reach set over-approximation for nonlinear systems. In 2016 International Conference on Embedded Software (EMSOFT), pages 1–10, Pittsburgh, PA, USA, 2016. C. Fan, J. Kapinski, X. Jin, and S. Mitra. Locally optimal reach set over-approximation for nonlinear systems. In 2016 International Conference on Embedded Software (EMSOFT), pages 1–10, Pittsburgh, PA, USA, 2016.
21.
Zurück zum Zitat C. Fan, B. Qi, S. Mitra, and M. Viswanathan. DryVR: Data-driven verification and compositional reasoning for automotive systems. In Computer Aided Verification CAV 2017, Heidelberg, Germany, July 24-28, 2017, Proceedings, Part I, volume 10426 of LNCS, pages 441–461. Springer, 2017. C. Fan, B. Qi, S. Mitra, and M. Viswanathan. DryVR: Data-driven verification and compositional reasoning for automotive systems. In Computer Aided Verification CAV 2017, Heidelberg, Germany, July 24-28, 2017, Proceedings, Part I, volume 10426 of LNCS, pages 441–461. Springer, 2017.
22.
Zurück zum Zitat J. Fan, C. Huang, X. Chen, W. Li, and Q. Zhu. ReachNN*: A tool for reachability analysis of neural-network controlled systems. In International Symposium on Automated Technology for Verification and Analysis, pages 537–542. Springer, 2020. J. Fan, C. Huang, X. Chen, W. Li, and Q. Zhu. ReachNN*: A tool for reachability analysis of neural-network controlled systems. In International Symposium on Automated Technology for Verification and Analysis, pages 537–542. Springer, 2020.
23.
Zurück zum Zitat G. Frehse, C. L. Guernic, A. Donzé, S. Cotton, R. T. Iosif, R. Passerone, and T. Dang. Spaceex: Scalable verification of hybrid systems. In G. Gopalakrishnan and S. Qadeer, editors, Computer Aided Verification, CAV 2011, volume 6806 of Lecture Notes in Computer Science, pages 379–395. Springer, Berlin, Heidelberg, 2011. G. Frehse, C. L. Guernic, A. Donzé, S. Cotton, R. T. Iosif, R. Passerone, and T. Dang. Spaceex: Scalable verification of hybrid systems. In G. Gopalakrishnan and S. Qadeer, editors, Computer Aided Verification, CAV 2011, volume 6806 of Lecture Notes in Computer Science, pages 379–395. Springer, Berlin, Heidelberg, 2011.
24.
Zurück zum Zitat L. Geretti, J. A. D. Sandretto, M. Althoff, L. Benet, P. Collins, M. Forets, E. Ivanova, Y. Li, S. Mitra, S. Mitsch, C. Schilling, M. Wetzlinger, and D. Zhuang. Arch-comp23 category report: Continuous and hybrid systems with nonlinear dynamics. In G. Frehse and M. Althoff, editors, Proceedings of 10th International Workshop on Applied Verification of Continuous and Hybrid Systems (ARCH23), volume 96 of EPiC Series in Computing, pages 61–88. EasyChair, 2023. L. Geretti, J. A. D. Sandretto, M. Althoff, L. Benet, P. Collins, M. Forets, E. Ivanova, Y. Li, S. Mitra, S. Mitsch, C. Schilling, M. Wetzlinger, and D. Zhuang. Arch-comp23 category report: Continuous and hybrid systems with nonlinear dynamics. In G. Frehse and M. Althoff, editors, Proceedings of 10th International Workshop on Applied Verification of Continuous and Hybrid Systems (ARCH23), volume 96 of EPiC Series in Computing, pages 61–88. EasyChair, 2023.
25.
Zurück zum Zitat R. Goebel, R. G. Sanfelice, and A. R. Teel. Hybrid dynamical systems. IEEE Control Systems Magazine, 29(2):28–93, 2009. R. Goebel, R. G. Sanfelice, and A. R. Teel. Hybrid dynamical systems. IEEE Control Systems Magazine, 29(2):28–93, 2009.
26.
Zurück zum Zitat S. Gruenbacher, R. M. Hasani, M. Lechner, J. Cyranka, S. A. Smolka, and R. Grosu. On the verification of neural odes with stochastic guarantees. In Proceedings of the AAAI Conference on Artificial Intelligence, 2021. S. Gruenbacher, R. M. Hasani, M. Lechner, J. Cyranka, S. A. Smolka, and R. Grosu. On the verification of neural odes with stochastic guarantees. In Proceedings of the AAAI Conference on Artificial Intelligence, 2021.
27.
Zurück zum Zitat S. A. Gruenbacher, M. Lechner, R. Hasani, D. Rus, T. A. Henzinger, S. A. Smolka, and R. Grosu. Gotube: Scalable statistical verification of continuous-depth models. Proceedings of the AAAI Conference on Artificial Intelligence, 36(6):6755–6764, Jun. 2022. S. A. Gruenbacher, M. Lechner, R. Hasani, D. Rus, T. A. Henzinger, S. A. Smolka, and R. Grosu. Gotube: Scalable statistical verification of continuous-depth models. Proceedings of the AAAI Conference on Artificial Intelligence, 36(6):6755–6764, Jun. 2022.
28.
Zurück zum Zitat C. Huang, J. Fan, X. Chen, W. Li, and Q. Zhu. Polar: A polynomial arithmetic framework for verifying neural-network controlled systems. Proceedings of the AAAI Conference on Artificial Intelligence, 37(14):13220–13228, 2023. C. Huang, J. Fan, X. Chen, W. Li, and Q. Zhu. Polar: A polynomial arithmetic framework for verifying neural-network controlled systems. Proceedings of the AAAI Conference on Artificial Intelligence, 37(14):13220–13228, 2023.
29.
Zurück zum Zitat K. A. Khan and P. I. Barton. Generalized derivatives for solutions of parametric ordinary differential equations with non-differentiable right-hand sides. Journal of Optimization Theory and Applications, 163(2):355–386, 2014. K. A. Khan and P. I. Barton. Generalized derivatives for solutions of parametric ordinary differential equations with non-differentiable right-hand sides. Journal of Optimization Theory and Applications, 163(2):355–386, 2014.
30.
Zurück zum Zitat K. A. Khan and P. I. Barton. A vector forward mode of automatic differentiation for generalized derivative evaluation. Optimization Methods and Software, 30(6):1185–1212, 2015. K. A. Khan and P. I. Barton. A vector forward mode of automatic differentiation for generalized derivative evaluation. Optimization Methods and Software, 30(6):1185–1212, 2015.
31.
Zurück zum Zitat S. Kong, S. Gao, W. Chen, and E. Clarke. dreach: \(\delta \)-reachability analysis for hybrid systems. In Tools and Algorithms for the Construction and Analysis of Systems: 21st International Conference, TACAS 2015, Held as Part of the European Joint Conferences on Theory and Practice of Software, ETAPS 2015, London, UK, April 11-18, 2015, Proceedings 21, pages 200–205. Springer, 2015. S. Kong, S. Gao, W. Chen, and E. Clarke. dreach: \(\delta \)-reachability analysis for hybrid systems. In Tools and Algorithms for the Construction and Analysis of Systems: 21st International Conference, TACAS 2015, Held as Part of the European Joint Conferences on Theory and Practice of Software, ETAPS 2015, London, UK, April 11-18, 2015, Proceedings 21, pages 200–205. Springer, 2015.
32.
Zurück zum Zitat Y. Li, H. Zhu, K. Braught, K. Shen, and S. Mitra. Verse: A python library for reasoning about multi-agent hybrid system scenarios. In Computer Aided Verification - 35th International Conference, CAV 2023, Paris, France, July 17-22, 2023, Proceedings, Part I, pages 351–364, 2023. Y. Li, H. Zhu, K. Braught, K. Shen, and S. Mitra. Verse: A python library for reasoning about multi-agent hybrid system scenarios. In Computer Aided Verification - 35th International Conference, CAV 2023, Paris, France, July 17-22, 2023, Proceedings, Part I, pages 351–364, 2023.
33.
Zurück zum Zitat D. M. Lopez, P. Musau, N. P. Hamilton, and T. T. Johnson. Reachability analysis of a general class of neural ordinary differential equations. In S. Bogomolov and D. Parker, editors, Formal Modeling and Analysis of Timed Systems - 20th International Conference, FORMATS 2022, Warsaw, Poland, September 13-15, 2022, Proceedings, volume 13465 of Lecture Notes in Computer Science, pages 258–277. Springer, 2022. D. M. Lopez, P. Musau, N. P. Hamilton, and T. T. Johnson. Reachability analysis of a general class of neural ordinary differential equations. In S. Bogomolov and D. Parker, editors, Formal Modeling and Analysis of Timed Systems - 20th International Conference, FORMATS 2022, Warsaw, Poland, September 13-15, 2022, Proceedings, volume 13465 of Lecture Notes in Computer Science, pages 258–277. Springer, 2022.
34.
Zurück zum Zitat J. Maidens and M. Arcak. Reachability analysis of nonlinear systems using matrix measures. Automatic Control, IEEE Transactions on, 60(1):265–270, 2015. J. Maidens and M. Arcak. Reachability analysis of nonlinear systems using matrix measures. Automatic Control, IEEE Transactions on, 60(1):265–270, 2015.
35.
Zurück zum Zitat S. Mitra. Verifying cyber-physical systems: A path to safe autonomy. The MIT Press, 2021. S. Mitra. Verifying cyber-physical systems: A path to safe autonomy. The MIT Press, 2021.
36.
Zurück zum Zitat Y. Nesterov. Lexicographic differentiation of nonsmooth functions. Mathematical Programming, 104(2-3):669–700, 2005. Y. Nesterov. Lexicographic differentiation of nonsmooth functions. Mathematical Programming, 104(2-3):669–700, 2005.
37.
Zurück zum Zitat O. Ogunmolu, X. Gu, S. Jiang, and N. Gans. Nonlinear systems identification using deep dynamic neural networks. arXiv preprint arXiv:1610.01439, 2016. O. Ogunmolu, X. Gu, S. Jiang, and N. Gans. Nonlinear systems identification using deep dynamic neural networks. arXiv preprint arXiv:​1610.​01439, 2016.
38.
Zurück zum Zitat G. Pillonetto, A. Aravkin, D. Gedon, L. Ljung, A. H. Ribeiro, and T. B. Schön. Deep networks for system identification: a survey. arXiv preprint arXiv:2301.12832, 2021. G. Pillonetto, A. Aravkin, D. Gedon, L. Ljung, A. H. Ribeiro, and T. B. Schön. Deep networks for system identification: a survey. arXiv preprint arXiv:​2301.​12832, 2021.
39.
Zurück zum Zitat H. Roehm, J. Oehlerking, T. Heinz, and M. Althoff. Stl model checking of continuous and hybrid systems. In Proceedings of the International Conference on Tools and Algorithms for the Construction and Analysis of Systems (TACAS), pages 468–486. Springer, 2016. H. Roehm, J. Oehlerking, T. Heinz, and M. Althoff. Stl model checking of continuous and hybrid systems. In Proceedings of the International Conference on Tools and Algorithms for the Construction and Analysis of Systems (TACAS), pages 468–486. Springer, 2016.
40.
Zurück zum Zitat H. Royden and P. Fitzpatrick. Real Analysis. Prentice Hall, Upper Saddle River, NJ, 4th edition, 2010. H. Royden and P. Fitzpatrick. Real Analysis. Prentice Hall, Upper Saddle River, NJ, 4th edition, 2010.
41.
Zurück zum Zitat Z. Shi, Y. Wang, H. Zhang, J. Z. Kolter, and C.-J. Hsieh. Efficiently computing local lipschitz constants of neural networks via bound propagation. CoRR, abs/2210.07394, 2022. Z. Shi, Y. Wang, H. Zhang, J. Z. Kolter, and C.-J. Hsieh. Efficiently computing local lipschitz constants of neural networks via bound propagation. CoRR, abs/2210.07394, 2022.
42.
Zurück zum Zitat H.-D. Tran, F. Cai, M. D. Lopez, P. Musau, T. T. Johnson, and X. Koutsoukos. Safety verification of cyber-physical systems with reinforcement learning control. ACM Transactions on Embedded Computing Systems (TECS), 18(5s), October 2019. H.-D. Tran, F. Cai, M. D. Lopez, P. Musau, T. T. Johnson, and X. Koutsoukos. Safety verification of cyber-physical systems with reinforcement learning control. ACM Transactions on Embedded Computing Systems (TECS), 18(5s), October 2019.
43.
Zurück zum Zitat H. D. Tran, X. Yang, D. M. Lopez, P. Musau, L. V. Nguyen, W. Xiang, S. Bak, and T. T. Johnson. NNV: The neural network verification tool for deep neural networks and learning-enabled cyber-physical systems. In Proceedings of the 32nd International Conference on Computer-Aided Verification (CAV), July 2020. H. D. Tran, X. Yang, D. M. Lopez, P. Musau, L. V. Nguyen, W. Xiang, S. Bak, and T. T. Johnson. NNV: The neural network verification tool for deep neural networks and learning-enabled cyber-physical systems. In Proceedings of the 32nd International Conference on Computer-Aided Verification (CAV), July 2020.
44.
Zurück zum Zitat R.-T. Wu and M. R. Jahanshahi. Deep convolutional neural networks for structural dynamic response estimation and system identification. Journal of Engineering Mechanics, 145(1), 2019. R.-T. Wu and M. R. Jahanshahi. Deep convolutional neural networks for structural dynamic response estimation and system identification. Journal of Engineering Mechanics, 145(1), 2019.
45.
Zurück zum Zitat K. Xu, Z. Shi, H. Zhang, Y. Wang, K.-W. Chang, M. Huang, B. Kailkhura, X. Lin, and C.-J. Hsieh. Automatic perturbation analysis for scalable certified robustness and beyond. Advances in Neural Information Processing Systems, 33:1129–1141, 2020. K. Xu, Z. Shi, H. Zhang, Y. Wang, K.-W. Chang, M. Huang, B. Kailkhura, X. Lin, and C.-J. Hsieh. Automatic perturbation analysis for scalable certified robustness and beyond. Advances in Neural Information Processing Systems, 33:1129–1141, 2020.
46.
Zurück zum Zitat H. Zhang, T.-W. Weng, P.-Y. Chen, C.-J. Hsieh, and L. Daniel. Efficient neural network robustness certification with general activation functions. In Advances in Neural Information Processing Systems, pages 4939–4948, 2018. H. Zhang, T.-W. Weng, P.-Y. Chen, C.-J. Hsieh, and L. Daniel. Efficient neural network robustness certification with general activation functions. In Advances in Neural Information Processing Systems, pages 4939–4948, 2018.
47.
Zurück zum Zitat Q. Zhao, X. Chen, Z. Zhao, Y. Zhang, E. Tang, and X. Li. Verifying neural network controlled systems using neural networks. In Proceedings of the 25th ACM International Conference on Hybrid Systems: Computation and Control, HSCC ’22, New York, NY, USA, 2022. Association for Computing Machinery. Q. Zhao, X. Chen, Z. Zhao, Y. Zhang, E. Tang, and X. Li. Verifying neural network controlled systems using neural networks. In Proceedings of the 25th ACM International Conference on Hybrid Systems: Computation and Control, HSCC ’22, New York, NY, USA, 2022. Association for Computing Machinery.
Metadaten
Titel
Reachability for Nonsmooth Systems with Lexicographic Jacobians
verfasst von
Chenxi Ji
Huan Zhang
Sayan Mitra
Copyright-Jahr
2025
DOI
https://doi.org/10.1007/978-3-031-90653-4_18