Discontinuous_dynamical_systems
Discontinuous_dynamical_systems
NONSMOOTH ANALYSIS,
AND STABILITY
JORGE CORTÉS
© PHOTOCREDIT
iscontinuous dynamical systems arise in a large manipulation of objects by means of mechanical contact
Example 4: Move-Away-from-
Nearest-Neighbor Interaction Law
Consider n nodes p1 , . . . , pn evolving in a square Q
according to the interaction rule “move diametrically
away from the nearest neighbor.” Note that this rule is not
defined when two nodes are located at the same point,
that is, when the configuration is an element of S
{(p1 , . . . , pn ) ∈ Qn : pi = pj for some i = j} . We thus
consider only configurations that belong to Qn \S . Next,
we define the map that assigns to each node its nearest
F υ
neighbor, where the nearest neighbor may be an element
of the boundary bndry(Q) of the square. Note that the
mg nearest neighbor of a node might not be unique, that is,
θ more than one node can be located at the same (nearest)
distance. Hence, we define the nearest-neighbor map
N = (N1 , . . . , Nn ) : Qn \ S → Qn by arbitrarily selecting,
(a)
for each i ∈ {1, . . . , n}, an element,
2
Ni(p1 , . . . , pn ) ∈ argmin{||pi − q||2
1.5 : q ∈ bndry(Q) ∪ {p1 , . . . , pn } \ {pi}},
ν 1
where argmin stands for the minimizing value of q, and
0.5
|| · ||2 denotes the Euclidean norm. By definition,
Ni(p1 , . . . , pn ) = pi . For i ∈ {1, . . . , n}, we thus define the
0 move-away-from-nearest-neighbor interaction law
−2 −1 0 1 2
υ pi(t) − Ni(p1 (t), . . . , pn (t))
(b) ṗi = . (6)
||pi(t) − Ni(p1 (t), . . . , pn (t))||2
FIGURE 2 Brick sliding on a frictional ramp with velocity ν. (a) The Changes in the value of the nearest-neighbor map N induce
physical quantities used to describe the example and (b) the one- discontinuities in the dynamical system. For instance, con-
dimensional phase portraits of (2) corresponding to values of the fric-
tion coefficient ν between zero and two, with a ramp inclination of
sider a node sufficiently close to a vertex of bndry(Q) so
30◦ . The phase portrait shows that, for sufficiently small values of ν, that the closest neighbor to the node is an element of the
every trajectory that starts with positive initial velocity (“moving to the boundary. Depending on how the node is positioned with
right”) never stops. However, for sufficiently large values of ν, every respect to the bisector line passing through the vertex, the
trajectory that starts with positive initial velocity eventually stops and node computes different directions of motion; see Figure 4
remains stopped. However, there is no continuously differentiable
solution of (2) that exhibits this type of behavior. We thus need to
for an illustration.
expand our notion of solution beyond continuously differentiable To analyze the dynamical system (6), we need to
solutions by understanding the effect of the discontinuity in (2). understand how the discontinuities affect its evolution.
.5 .5
x2
0
x2
−.5 −.5
−1
−1
−1 −.5 0 .5 1 −1 −.5 0 .5 1
x1 x1
(a) (b)
FIGURE 3 Nonsmooth harmonic oscillator. (a) The phase portrait on [−1, 1]2 of the vector field (x1 , x 2 ) → (x 2 , −sign (x1 )) and (b) the
contour plot on [−1, 1]2 of the function (x1 , x 2 ) → |x1 | + (x 22 /2). The discontinuity of the vector field along the x 2 -coordinate axis
makes it impossible to find continuously differentiable solutions of (10). On the other hand, the level sets in (b) match the phase por-
trait in (a) everywhere except for the x 2 -coordinate axis, which suggests that trajectories along the level sets are candidates for solu-
tions of (10) in a sense different from the classical one.
Since each node moves away from its nearest neighbors, The function γ : [a, b] → R is absolutely continuous if, for
it is reasonable to expect that the nodes never run into all ∈ (0, ∞), there exists δ ∈ (0, ∞) such that, for each
each other, however, rigorous verification of this proper- finite collection {(a1 , b1 ), . . . , (an , bn )} of disjoint open
ty requires a proof. We would also like to characterize intervals contained in [a, b] with ni=1 (bi − ai) < δ, it fol-
the asymptotic behavior of the trajectories of the system lows that
(6). To study these questions, we need to extend our
n
notion of solution. |γ (bi) − γ (ai)| < .
i=1
Beyond Continuously Differentiable Solutions
Examples 2–4 can be described by a dynamical system of Equivalently [12], γ is absolutely continuous if there exists
the form a Lebesgue integrable function κ : [a, b] → R such that
I
n this article, we focus entirely on absolutely continuous solutions [S3] R.W. Cottle, J.-S. Pang, and R.E. Stone, The Linear Comple-
of ordinary differential equations. However, nonsmooth continu- mentarity Problem. New York: Academic, 1992.
ous-time systems can possess discontinuous solutions that admit [S4] G. Isac, Complementarity Problems. New York: Springer-Ver-
jumps in the state. Such notions are appropriate for dealing with lag, 1992.
mechanical systems, for instance, subject to unilateral constraints [S5] M.C. Ferris and J.-S. Pang, “Engineering and economic applica-
[15]. As an example, a bouncing ball hitting the ground experiences tions of complementarity problems,” SIAM Rev., vol. 39, no. 4, pp.
an instantaneous change of velocity. This change corresponds to a 669–713, 1997.
discontinuous jump in the trajectory describing the evolution of the [S6] B. Brogliato, “Some perspectives on the analysis and control of
velocity of the ball. Both measure differential inclusions [S1], [S2] complementarity systems,” IEEE Trans. Automat. Contr., vol. 48, no.
and linear complementarity systems [S3]–[S6] are approaches that 6, pp. 918–935, 2003.
specifically allow for discontinuous solutions. Within hybrid systems [S7] P.J. Antsaklis and A. Nerode, guest eds., “Special issue on
theory [S7]–[S10], discontinuous solutions arise in systems that hybrid control systems,” IEEE Trans. Automat. Contr., vol. 43, no. 4,
involve both continuous- and discrete-time evolutions. 1998.
[S8] A.S. Morse, C.C. Pantelides, S.S. Sastry, and J.M. Schu-
REFERENCES macher, Guest Eds., “Special issue on hybrid systems,” Automat-
[S1] J.-J. Moreau, “Unilateral contact and dry friction in finite free- ica, vol. 35, no. 3, 1999.
dom dynamics,” in Non-Smooth Mechanics and Applications, vol. [S9] A.J. van der Schaft and H. Schumacher, An Introduc-
302, J.-J. Moreau and P.D. Panagiotopoulos, Eds. New York: tion to Hybrid Dynamical Systems. New York: Springer-Ver-
Springer-Verlag, 1988, pp. 1–82. lag, 2000.
[S2] M.D. P.M. Marques, Differential Inclusions in Nonsmooth [S10] R. Goebel and A.R. Teel, “Solutions of hybrid inclusions via set
Mechanical Problems: Shocks and Dry Friction. Cambridge, MA: and graphical convergence with stability theory applications,” Auto-
Birkhäuser, 1993. matica, vol. 42, no. 4, pp. 573–587, 2006.
S
everal solution notions are available in addition to Caratheodory, [S14] V.A. Yakubovich, G.A. Leonov, and A.K. Gelig, in Stability of
Filippov, and sample-and-hold solutions. These notions include Stationary Sets in Control Systems With Discontinuous Nonlineari-
the ones considered by Krasovskii [34], Hermes [S11], [S12], ties, vol. 14. Singapore: World Scientific, 2004.
Ambrosio [S13], Sentis [35], and Yakubovich-Leonov-Gelig [S14]; [S15] J.S. Spraker and D.C. Biles, “A comparison of the
see Table S1. As in the case in which the vector field is continuous, Caratheodory and Filippov solution sets,” J. Math. Anal. Appl., vol.
Euler solutions [18], [24] are useful for establishing existence and in 198, no. 2, pp. 571–580, 1996.
characterizing basic mathematical properties of the dynamical sys- [S16] S. Hu, “Differential equations with discontinuous right-hand
tem. Additional notions of solutions for discontinuous systems are sides,” J. Math. Anal. Appl., vol. 154, no. 2, pp. 377–390, 1991.
[S17] A. Bacciotti, “Some remarks on generalized solutions of dis-
provided in [S14, sect.1.1.3]. With so many notions of solution avail-
continuous differential equations,” Int. J. Pure Applied Math., vol. 10,
able, various works explore the relationships among them. For
no. 3, pp. 257–266, 2004.
example, Caratheodory and Filippov solutions are compared in
[S15]; Caratheodory and Krasovskii solutions are compared in
[S16]; Caratheodory, Euler, sample-and-hold, Filippov, and
TABLE S1 Several notions of solution for discontinuous
Krasovskii solutions are compared in [16]; Hermes, Filippov, and dynamics. Depending on the specific problem, some
Krasovskii solutions are compared in [S12]; and Caratheodory, notions give more physically meaningful solution
Euler, and Sentis solutions are compared in [S17]. trajectories than others.
B
eyond the topics discussed in this article, we briefly mention 37, pp. 813–840, 1999.
several that are relevant to systems and control. These top- [S19] F.H. Clarke, Y.S. Ledyaev, L. Rifford, and R.J. Stern, “Feed-
ics include continuous dependence of solutions with respect to back stabilization and Lyapunov functions,” SIAM J. Control Optim.,
initial conditions and parameters [18], [19], robustness proper- vol. 39, no. 1, pp. 25–48, 2000.
ties against external disturbances and state measurement [S20] A.R. Teel and L. Praly, “A smooth Lyapunov function from a class
errors [24], [S18], conditions for the existence of periodic solu- KL estimate involving two positive semidefinite functions,” ESAIM J. Con-
tions [18], [20], bifurcations [38], converse Lyapunov theorems trol, Optim. Calculus Variations, vol. 5, pp. 313–367, 2000.
[S19], [S20], controllability of differential inclusions [20], [24], [S21] J.P. Aubin, Viability Theory. Cambridge, MA: Birkhäuser, 1991.
output tracking in differential inclusions [43], systems subject to [S22] A. Nagurney and D. Zhang, Projected Dynamical Systems
gradient nonlinearities [S14], viability theory [S21], maximal and Variational Inequalities with Applications. Norwell, MA: Kluwer,
monotone inclusions [19], projected dynamical systems and 1996.
variational inequalities [S22], and numerical methods for dis- [S23] A. Dontchev and F. Lempio, “Difference methods for differential
continuous systems and differential inclusions [S23]–[S25]. inclusions: A survey,” SIAM Rev., vol. 34, pp. 263–294, 1992.
Several works report equivalence results among different [S24] F. Lempio and V. Veliov, “Discrete approximations of differential
approaches to nonsmooth systems, including [19] on the equiv- inclusions,” Bayreuth. Math. Schr., vol. 54, pp. 149–232, 1998.
alence between differential inclusions and projected dynamical [S25] V. Acary and B. Brogliato, Numerical Methods for Nonsmooth
systems and [S26] on the equivalence of these formalisms with Dynamical Systems: Applications in Mechanics and Electronics.
complementarity systems. New York: Springer-Verlag, 2008.
[S26] B. Brogliato, A. Daniilidis, C. Lemaréchal, and V. Acary, “On
REFERENCES the equivalence between complementarity systems, projected sys-
[S18] Y.S. Ledyaev and E.D. Sontag, “A Lyapunov characterization tems, and differential inclusions,” Syst. Control Lett., vol. 55, no. 1,
of robust stabilization,” Nonlinear Anal., Theory, Methods, Appl., vol. pp. 45–51, 2006.
Index of Symbols
G [X ] Set-valued map associated with a control system ne Unit normal to the edge e of a polygon Q pointing
x : Rd × U → R d toward the interior of Q
co(S ) Convex closure of a set S ⊆ Rd f Set of points where the locally Lipschitz function
co(S ) Convex hull of a set S ⊆ Rd f : Rd → R fails to be differentiable
diam(π) Diameter of the partition π π Partition of a closed interval
f o(x; v ) Generalized directional derivative of the function B(S) Set whose elements are all of the possible sub-
f : Rd → R at x ∈ Rd in the direction of v ∈ Rd sets of S ⊆ Rd
f (x; v ) Right directional derivative of the function ∂P f Proximal subdifferential of the lower semicontinu-
f : Rd → R at x ∈ Rd in the direction of v ∈ Rd ous function f : Rd → R
SX Set of points where the vector field x : Rd → Rd is L̃F f Set-valued Lie derivative of the locally Lipschitz
discontinuous function f : Rd → R with respect to the set-val-
dist ( p, S ) Euclidean distance from the point p ∈ Rd to the ued map F : Rd → B( Rd )
set S ⊆ Rd L̃ X f Set-valued Lie derivative of the locally Lipschitz
F [X ] Filippov set-valued map associated with a vector function f : Rd → R with respect to the Filippov
field X : Rd → Rd set-valued map F [X ] : Rd → B( Rd )
∂f Generalized gradient of the locally Lipschitz func- LF f Lower set-valued Lie derivative of the lower
tion f : Rd → R semicontinuous function f : Rd → R with
∇f Gradient of the differentiable function f : Rd → respect to the set-valued map F : Rd → B( Rd )
R LF f Upper set-valued Lie derivative of the lower
Ln(S ) Least-norm elements in the closure of the set semicontinuous function f : Rd → R with
S ⊆ Rd respect to the set-valued map F : Rd → B( Rd )
(x ) Set of limit points of a curve t → x(t) F Set-valued map
N Nearest-neighbor map sm Q Minimum distance function from a point in a con-
vex polygon Q ⊂ Rd to the boundary of Q
A
differential inclusion [19], [20] is a generalization of a dif- [24]. The uniqueness of Caratheodory solutions is guaranteed by
ferential equation. At each state, a differential inclusion the following result.
specifies a set of possible evolutions, rather than a single
one. This object is defined by means of a set-valued map; PROPOSITION S3
see “Set-Valued Maps.” The differential inclusion associated In addition to the hypotheses of Proposition S2, assume that, for all
with a time-varying set-valued map F : [0, ∞)× Rd → B(Rd) x ∈ Rd , there exist ε ∈ (0, ∞) and an integrable function L x :
is an equation of the form R → (0, ∞) such that
The point xe ∈ Rd is an equilibrium of the differential inclusion if for almost every y, y ∈ B(x, ε) , every t ∈ [0, ∞) , every
0 ∈ F(t, xe) for all t ∈ [0, ∞). We now define the notion of solution v ∈ F(t, y), and every w ∈ F(t, y ). Then, for all (t0 , x0 ) ∈ [0, ∞)×
of a differential inclusion in the sense of Caratheodory. Rd , there exists a unique Caratheodory solution of (S2) with initial
A Caratheodory solution of (S2) defined on [t0 , t1 ] ⊂ [0, ∞) is condition x(t0 ) = x0 .
an absolutely continuous map x : [t0 , t1 ] → Rd such that The following example illustrates propositions S2 and S3.
ẋ(t) ∈ F(t, x(t)) for almost every t ∈ [t0 , t1 ]. The existence of at Following [43], consider the set-valued map F : R → B(R)
least one solution starting from each initial condition is guaran- defined by
teed by the following result (see, for instance, [14] and [19]).
0, x = 0,
F(x) =
[−1, 1], x = 0.
PROPOSITION S2
Let F : [0, ∞)× Rd →B (Rd) be locally bounded and take non- Note that F is upper semicontinuous, but not lower semicontinu-
empty, compact, and convex values. Assume that, for each t ∈ R, ous, and thus it is not continuous. This set-valued map satisfies all
the set-valued map x → F(t, x) is upper semicontinuous, and, for of the hypotheses in Proposition S2, and therefore Caratheodory
each x ∈ Rd , the set-valued map t → F(t, x) is measurable. Then, solutions exist starting from all initial conditions. In addition, F sat-
for all (t0 , x 0 ) ∈ [0, ∞)× Rd , there exists a Caratheodory solution of isfies (S3) as long as y and y are nonzero. Therefore, Proposi-
(S1) with initial condition x(t0 ) = x 0 . tion S3 guarantees the uniqueness of Caratheodory solutions. In
This result is sufficient for our purposes. Additional existence fact, for every initial condition, the Caratheodory solution of
results based on alternative assumptions are given in [14] and ẋ(t ) ∈ F(x(t )) is just the equilibrium solution.
1 1
.5 .5
−1 −.5 .5 1 −1 −.5 .5 1
−.5 −.5
−1 −1
(a) (b)
1
.3
.8
.2
.6 .1
.4
−1 −.5 .5 1
.2 −.1
−.2
−1 −.5 .5 1
(c) (d)
FIGURE 5 Discontinuous and not-one-sided Lipschitz vector fields. The vector fields in (a) and (b), which differ only in their values at zero,
are discontinuous, and thus do not satisfy the hypotheses of Proposition 1. Therefore, the existence of solutions is not guaranteed. In fact,
the vector field in (a) has no solution starting from zero, whereas the vector field in (b) has a solution starting from all initial conditions. The
vector fields in (c) and (d) are neither locally Lipschitz nor one-sided Lipschitz, and thus do not satisfy the hypotheses of Proposition 2.
Therefore, uniqueness of solutions is not guaranteed. The vector field in (c) has two solutions starting from zero. However, the vector field
in (d) has a unique solution starting from all initial conditions.
.6
f (y) − f (y )2 ≤ Lx y − y 2 ,
.4
xn − x X(x)
− < δ, n ∈ N, (18)
xn − x2 X(x)2 2
x2
0
it follows that X(xn ) → X(x). If the vector field X is direction-
ally continuous, then, for all x0 ∈ Rd , there exists a
−.5 Caratheodory solution of (10) with initial condition x(0) = x0 .
Patchy vector fields [30] are another class of time-invariant,
discontinuous vector fields that have Caratheodory solutions.
Additional conditions for the existence and uniqueness of
−1
−1 −.5 0 .5 1
Caratheodory conditions can be found in [31]–[33]. Beyond dif-
x1 ferential equations, Caratheodory solutions can also be defined
(a) for differential inclusions, as explained in “Set-Valued Maps”
and “Caratheodory Solutions of Differential Inclusions.”
1 Filippov Solutions
The above discussion shows that the relaxation of the value
of the vector field on a set of times of measure zero in the def-
.5 inition of Caratheodory solution is not always sufficient to
guarantee that such solutions exist. Due to the discontinuity
of the vector field, its value can exhibit significant variations
arbitrarily close to a given point, and this mismatch might
0
x2
a − 12 sign(a)t, a − 12 sign(a)t , t ≤ |a|, F[X](x) = {X(x)}. (25)
t → x(t) =
(0, 0), t ≥ |a|.
Sum Rule: If X1 , X2 : Rd → Rm are locally bounded at
(23) x ∈ Rd , then
Note that the solution slides along the diagonals, following a F[X1 + X2 ](x) ⊆ F[X1 ](x) + F[X2 ](x). (26)
convex combination of the limiting values of X around the
diagonals, rather than the direction specified by X itself. We Moreover, if either X1 or X2 is continuous at x, then equal-
study this type of behavior in more detail in the section “Piece- ity holds.
wise Continuous Vector Fields and Sliding Motions.” Product Rule: If X1 : Rd → Rm and X2 : Rd → Rn are
locally bounded at x ∈ Rd , then
Relationship Between Caratheodory and Filippov Solutions
In general, Caratheodory and Filippov solutions are not
related. A vector field for which both notions of solution F[(X1 , X2 )T ](x) ⊆ F[X1 ](x) × F[X2 ](x). (27)
⎧
⎪ {(y1 , y2 ) ∈ R2 : |y1 + y2 | ≤ 1, |y1 − y2 | ≤ 1}, (x1 , x2 ) = (0, 0),
⎪
⎪
⎪
⎪ {(−1, 0)}, −x1 < x2 < x1 ,
⎪
⎪
⎪ {(y1 , y2 ) ∈ R2
⎪ : y1 + y2 = −1, y1 ∈ [−1, 0]}, 0 < x2 = x1 ,
⎪
⎪
⎪
⎪
⎨ {(0, 1)}, x2 < x1 < −x2 ,
F [X](x1 , x2 ) = {(y1 , y2 ) ∈ R2 : y1 − y2 = −1, y1 ∈ [−1, 0], 0 < −x1 = x2 , (22)
⎪
⎪
⎪
⎪ {(1, 0)}, x1 < x2 < −x1 ,
⎪
⎪
⎪
⎪ 1 , y2 ) ∈ R2
{(y : y1 + y2 = 1, y1 ∈ [0, 1]}, x2 = x1 < 0,
⎪
⎪
⎪
⎪ {(0, −1)}, −x2 < x1 < x2 ,
⎪
⎩
{(y1 , y2 ) ∈ R2 : y1 − y2 = 1, y1 ∈ [0, 1]}, 0 < x1 = −x2 .
This set-valued map can be computed as follows. At points D1 = {(x1 , x2 ) ∈ [−1, 1]2 : −x1 < x2 < x1 },
of continuity of X, that is, for x ∈
/ SX , the consistency prop-
D2 = {(x1 , x2 ) ∈ [−1, 1]2 : x2 < x1 < −x2 },
erty (25) implies that F[X](x) = {X(x)}. At points of discon-
tinuity of X, that is, for x ∈ SX , F[X](x) is a convex D3 = {(x1 , x2 ) ∈ [−1, 1]2 : x1 < x2 < −x1 },
polyhedron in Rd of the form D4 = {(x1 , x2 ) ∈ [−1, 1]2 : −x2 < x1 < x2 }.
F [X](x) = co XD (x) : x ∈ bndry(Dk ) . (31) Its Filippov set-valued map, described in (22), maps points
k
outside the diagonals SX = {(a, ±a) ∈ [−1, 1]2 : a ∈ [−1, 1]}
As an illustration, let us revisit examples 2–4. to singletons, maps points in SX \{(0, 0)} to closed seg-
ments, and maps (0, 0) to a square polygon. Several Filip-
Examples 2–4 Revisited: Computation of the Filippov pov solutions starting from various initial conditions are
Set-Valued Map and Filippov Solutions plotted in Figure 6(b).
The vector field of the sliding brick in Example 2 is Let us now discuss what happens on the points of dis-
piecewise continuous with D1 = {v ∈ R : v < 0} and continuity of the vector field X : Rd → Rd . Suppose that
D2 = {v ∈ R : v > 0}. Note that the restriction of the vec- x ∈ SX belongs to just two boundary sets, that is,
tor field to D1 can be continuously extended to D1 by set- x ∈ bndry (Di) ∩ bndry (D j) , for some distinct i, j ∈
ting X|D (0) = g(sin θ + ν cos θ). Likewise, the restriction {1, . . . , m}, but x ∈
/ bndry (Dk ), for all k ∈ {1, . . . , m}\{i, j}.
1
at every point in SX ∩ {x2 < 0}; see Figure 3(a). Moreover, Solutions by Means of Differential Inclusions
there is only one solution (the equilibrium solution) start- A first alternative to defining a solution notion consists of
ing from (0, 0). Therefore, using Proposition 5, we con- associating a differential inclusion with the control equa-
clude that this system has a unique Filippov solution tion (33). In this approach, the set-valued map
starting from each initial condition. G[X] : Rd → B (Rd) is defined by
For Example 4, it is convenient to define D5 = D1 . Then,
at (x1 , x2 ) ∈ bndry (Di) ∩ bndry (Di+1 ) \{(0, 0)}, with G[X](x) {X(x, u) : u ∈ U}. (34)
i ∈ {1, . . ., 4}, the vector X|D (x1 , x2 ) points into Di+1 , and
i
the vector X|D (x1 , x2 ) points into Di ; see Figure 6(a). In other words, the set G[X](x) captures all of the direc-
i+1
Moreover, there is only one solution (the equilibrium solu- tions in Rd that can be generated at x with controls belong-
tion) starting from (0, 0). Therefore, using Proposition 5, ing to U . Consider now the differential inclusion
we conclude that Example 4 has a unique Filippov solution
starting from each initial condition. ẋ(t) ∈ G[X](x(t)). (35)
Proposition 5 can also be applied to piecewise contin-
uous vector fields with an arbitrary number of partition- A solution of (33) on [0, t1 ] ⊂ R is defined to be a
ing domains, provided that the set where the vector Caratheodory solution of the differential inclusion (35),
field is discontinuous is a disjoint union of surfaces that is, an absolutely continuous map x : [0, t1 ] → Rd such
resulting from pairwise intersections of the boundaries that ẋ(t) ∈ G[X](x(t)) for almost all t ∈ [0, t1 ].
of pairs of domains. Alternative versions of this result If we choose an open-loop input u : [0, ∞) → U in
can also be stated for time-varying piecewise continuous (33), then a Caratheodory solution of the resulting
vector fields, as well as for situations in which more dynamical system is also a Caratheodory solution of the
than two domains intersect at a point of discontinuity differential inclusion (35). Alternatively, it can be
[18, thrm 4, p. 115]. shown [18] that, if X is continuous and U is compact,
The literature contains additional results guarantee- then the converse is also true. The differential inclusion
ing uniqueness of Filippov solutions tailored to specific (35) has the advantage of not focusing attention on a
classes of dynamical systems. For instance, [39] studies particular control input but rather allows us to compre-
uniqueness for relay linear systems, [11] investigates hensively study and understand the properties of the
uniqueness for adaptive control systems, while [40] control system.
establishes uniqueness for a class of discontinuous dif-
ferential equations whose vector field depends on the Sample-and-Hold Solutions
solution of a scalar conservation law. A second alternative to defining a solution notion for
the control equation (33) uses the notion of sample-
Solutions of Control Systems and-hold solution [41]. As discussed in the section
with Discontinuous Inputs “Stabilization of Control Systems,” this notion plays a
Let X : Rd × U → Rd , where U ⊆ Rm is the set of allowable key role in the stabilization of asymptotically control-
control-function values, and consider the control equation lable systems.
on Rd given by A partition of the interval [t0 , t1 ] is an increasing
sequence π = {si}N i =0 with s0 = t0 and sN = t1 . The
ẋ(t) = X(x(t), u(t)). (33) partition need not be finite. The notion of a partition of
[t0 , ∞) is defined similarly. The diameter of π is
At first sight, a natural way to identify a notion of solu- diam(π) sup{si − si−1 : i ∈ {1, . . . , N}} . Given a control
tion for (33) is to select a control input, either an open- input u : [0, ∞) × Rd → U , an initial condition x0 , and a
loop u : [0, ∞) → U , a closed-loop u : Rd → U , or a partition π of [0, t1 ], a π -solution of (33) defined on
combination u : [0, ∞) × Rd → U , and then consider the [0, t1 ] ⊂ R is the map x : [0, t1 ] → Rd , with x(0) = x0 ,
resulting differential equation. When the selected con- recursively defined by requiring the curve
trol input u is a discontinuous function of x ∈ Rd , then t ∈ [si−1 , si] → x(t) , for i ∈ {1, . . . , N − 1} , to be a
we can consider the solution notions of Caratheodory Caratheodory solution of the differential equation
or Filippov. At least two alternatives are considered in
the literature. We discuss them next. ẋ(t) = X(x(t), u(si−1 , x(si−1 ))). (36)
π -solutions are also referred to as sample-and-hold solutions The above example illustrates the need to consider non-
because the control is held fixed throughout each interval smooth analysis. Similar examples can be found in [14, sect.
of the partition at the value according to the state at the 2.2.2]. It is also worth noting that the presentation in “Stabili-
beginning of the interval. Figure 8 shows an example of a ty Analysis by Means of the Generalized Gradient of a Non-
π -solution. The existence of π -solutions is guaranteed [24] smooth Lyapunov Function” boils down to classical stability
if, for all u ∈ U ⊆ Rm , the map x → X(x, u) is continuous. analysis when the candidate Lyapunov function is smooth.
In this section we discuss two tools from nonsmooth
NONSMOOTH ANALYSIS analysis, namely, the generalized gradient and the proximal
We now consider candidate nonsmooth Lyapunov functions subdifferential [9], [24]. As with the notions of solution of
for discontinuous differential equations. The level of general- discontinuous differential equations, multiple generalized
ity provided by nonsmooth analysis is not always necessary. derivative notions are available in the literature when a func-
The stability properties of some discontinuous dynamical tion fails to be differentiable. These notions include, in addi-
systems and differential inclusions can be analyzed with tion to the two considered in this section, the generalized
smooth functions, as the following example shows. (super or sub) differential, the (upper or lower, right or left)
Dini derivative, and the contingent derivative [19], [24], [42],
Examples 5 and 6 Revisited: Stability [43]. Here, we focus on the notions of the generalized gradi-
of the Origin for the Sign Function ent and proximal subdifferential because of their role in pro-
We have already established that the vector fields (11) and viding stability tools for discontinuous differential equations.
(12) in examples 5 and 6, respectively, have unique Filip-
pov solutions starting from each initial condition. Now, The Generalized Gradient
consider the smooth Lyapunov function f : R → R, where of a Locally Lipschitz Function
f (x) = x2 /2. Now, for all x = 0, we have Rademacher’s theorem [9] states that every locally Lipschitz
function is differentiable almost everywhere in the sense of
∇ f (x) · X(x) = −|x| < 0. Lebesgue measure. When considering a locally Lipschitz
function as a candidate Lyapunov function, this statement
Since, according to (20), F[X](x) = {X(x)} on R \{0}, we may raise the question of whether to disregard those points
deduce that the function f is decreasing along every
Filippov solution of (11) and (12) that starts on R \{0}.
Therefore, we conclude that the equilibrium x = 0 is glob- x
ally asymptotically stable. 2.75
However, nonsmooth Lyapunov functions may be
needed if dealing with discontinuous dynamics, as the fol- 2.5
lowing example shows.
2.25
Example 3 Revisited: The Nonsmooth Harmonic Oscillator
2
Does Not Admit a Smooth Lyapunov Function
Consider the vector field for the nonsmooth harmonic
1.75
oscillator in Example 3. We reason as in [16]. As stated in
the section “Piecewise Continuous Vector Fields and Slid-
1.5
ing Motions,” all of the Filippov solutions of (3) are period-
ic and correspond to the curves plotted in Figure 3(b).
1.25
Therefore, if a smooth Lyapunov function f : R2 → R
exists, then it must be constant on each Filippov solution.
t
Since the level sets of f must be one dimensional, it fol- .2 .4 .6 .8 1
lows that each curve must be a level set. This property con-
FIGURE 8 Illustration of the notion of sample-and-hold solution. For
tradicts the fact that the function is smooth, since the level
the control system ẋ = u, we choose the control input u(x) = x. The
sets of a smooth function are also smooth, and the Filippov upper curve is the classical solution starting from x0 = 1, while the
solutions plotted in Figure 3(b) are not smooth along the lower curve is the sample-and-hold solution starting from x0 = 1
x2 -coordinate axis. corresponding to the π -partition {0, 1/4, 1/2, 3/4, 1} of [0, 1].
JUNE 2008 «
IEEE CONTROL SYSTEMS MAGAZINE 55
Authorized licensed use limited to: INDIAN INSTITUTE OF TECHNOLOGY ROORKEE. Downloaded on April 21,2025 at 03:11:55 UTC from IEEE Xplore. Restrictions apply.
where the gradient does not exist. Conceivably, the solu- everywhere except at zero, where it is only differentiable. It
tions of the dynamical system stay for almost all time away can be shown that ∇ f (0) = 0, while ∇ f (x) =
from “bad” points where no gradient of the function exists. 2x sin(1/x) − cos(1/x) for x = 0. From (37), we deduce that
However, this assumption is not always valid. As we show ∂ f (0) = [−1, 1], which is different from {∇ f (0)} = {0}. This
in Example 16 below, the solutions of a dynamical system example illustrates that Proposition 6-iii) may not be valid if
may insist on staying on the “bad” points forever. In that f is not continuously differentiable at the point x.
case, having some gradient-like information is helpful.
Let f : Rd → R be a locally Lipschitz function, and let Computing the Generalized Gradient
f ⊂ R denote the set of points where f fails to be
d Computation of the generalized gradient of a locally
differentiable. The generalized gradient ∂ f : Rd → B(Rd) of Lipschitz function is often a difficult task. In addition to
f is defined by the brute force approach of working from the definition
(37), various results are available to facilitate this compu-
tation. Many results that are valid for ordinary deriva-
∂ f (x) co lim ∇ f (xi) : xi → x , xi ∈
/ S∪ f , (37)
i→∞ tives have a counterpart in this setting. We summarize
some of these results here, and refer the reader to [9] and
where co denotes convex hull. In this definition, S ⊂ Rd is [24] for a complete exposition. In the statements of these
a set of measure zero that can be arbitrarily chosen to sim- results, the notion of a regular function plays a prominent
plify the computation. The resulting set ∂ f (x) is indepen- role; see “Regular Functions” for a precise definition.
dent of the choice of S. From the definition, the generalized Dilation Rule: If f : Rd → R is locally Lipschitz at x ∈ Rd
gradient of f at x consists of all convex combinations of all and s ∈ R, then the function s1 f1 + s2 f2 is locally Lipschitz
of the possible limits of the gradient at neighboring points at x, and
where f is differentiable. Equivalent definitions of the gen-
eralized gradient are given in [9]. ∂(sf )(x) = s∂ f (x). (38)
Some useful properties of the generalized gradient are
summarized in the following result [24]. Sum Rule: If f1 , f2 : Rd → R are locally Lipschitz at
x ∈ Rd and s1 , s2 ∈ R, then the function s1 f1 + s2 f2 is local-
Proposition 6 ly Lipschitz at x, and
If f : Rd → R is locally Lipschitz at x ∈ Rd , then the fol-
lowing statements hold: ∂(s1 f1 + s2 f2 )(x) ⊆ s1 ∂ f1 (x) + s2 ∂ f2 (x), (39)
i) ∂ f (x) is nonempty, compact, and convex.
ii) The set-valued map ∂ f : Rd → B(Rd), x → ∂ f (x), is where the sum of two sets A1 and A2 is defined as
upper semicontinuous and locally bounded at x. A1 + A2 = {a1 + a2 : a1 ∈ A1 , a2 ∈ A2 }. Moreover, if f1 and
iii) If f is continuously differentiable at x, then f2 are regular at x, and s1 , s2 ∈ [0, ∞), then equality holds
∂ f (x) = {∇ f (x)}. and s1 f1 + s2 f2 is regular at x.
Let us compute the generalized gradient for an illustra- Product Rule: If f1 , f2 : Rd → R are locally Lipschitz at
tive example. x ∈ Rd , then the function f1 f2 is locally Lipschitz at x, and
f (y + hv ) − f (y)
= lim+ sup .
Proposition 7 δ →0 y∈B(xδ)
ε →0+ h∈[0,ε)
h
For k ∈ {1, . . . , m}, let fk : Rd → R be locally Lipschitz at
x ∈ Rd , and define the functions fmax , fmin : Rd → R by The advantage of the generalized directional derivative compared
to the right directional derivative is that the limit always exists.
fmax (y) max{ fk (y) : k ∈ {1, . . . , m}}, When the right directional derivative exists, these quantities may
fmin (y) min{ fk (y) : k ∈ {1, . . . , m}}. be different. When they are equal, we call the function regular.
More formally, a function f : Rd → R is regular at x ∈ Rd if, for all
v ∈ Rd , the right directional derivative of f at x in the direction of v
Then, the following statements hold:
exists, and f (x; v) = f o(x; v). A function that is continuously dif-
i) fmax and fmin are locally Lipschitz at x.
ii) Let Imax (x) denote the set of indices k for which ferentiable at x is regular at x. Also, a convex function is regular
fk (x) = fmax (x). Then (cf. [9, Prop. 2.3.6]).
The function g : R → R, g (x ) = −|x |, is not regular. Since
g is continuously differentiable everywhere except for zero, it is
∂ fmax (x) ⊆ co {∂ fi(x) : i ∈ Imax (x)}. (43)
regular on R\{0}. However, its directional derivatives
Furthermore, if fi is regular at x for all i ∈ Imax (x), −v, v > 0, v, v > 0,
g (0; v) = go(0; v) =
v, v < 0, −v, v < 0,
then equality holds and fmax is regular at x.
iii) Let Imin (x) denote the set of indices k for which do not coincide. Hence, g is not regular at zero.
fk (x) = fmin (x). Then
∂ fmin (x) ⊆ co {∂ fi(x) : i ∈ Imin (x)}. (44)
⎧
⎨ {1}, x > 0,
Furthermore, if − fi is regular at x for all i ∈ Imin (x),
∂ f (x) = [−1, 1], x = 0, (45)
then equality holds and − fmin is regular at x. ⎩
{−1}, x < 0.
It follows from Proposition 7 that the maximum of a
finite set of continuously differentiable functions is a
locally Lipschitz and regular function whose generalized This result is obtained in Example 13 by direct
gradient at each point x is easily computable as the con- computation.
vex hull of the gradients of the functions that attain the
maximum at x.
Example 15: Generalized Gradient
Example 13 Revisited: Generalized Gradient of the Minus Absolute Value Function
of the Absolute Value Function The minimum of a finite set of regular functions is not
The absolute value function f (x) = |x| can be rewritten as always regular. A simple example is given by
f (x) = max{x, −x}. Both x → x and x → −x are continuous- g(x) = min{x, −x} = −|x|, which is not regular at zero, as
ly differentiable and hence locally Lipschitz and regular. we show in “Regular Functions.” However, according to
Therefore, according to Proposition 7-i) and -ii), f is locally Proposition 7-iii), this fact does not mean that its general-
Lipschitz and regular, and its generalized gradient is ized gradient cannot be computed. Indeed,
⎧
⎨ {−1}, x > 0, Example 16: Minimum-Distance-
∂g(x) = [−1, 1], x = 0, (46) to-Polygonal-Boundary Function
⎩
{1}, x < 0. Let Q ⊂ R2 be a convex polygon. Consider the minimum
distance function smQ : Q → R from a point within the
This result can also be obtained by combining (45) with the polygon to its boundary defined by
application of the dilation rule (38) to the function
f (x) = |x| with the parameter s = −1. smQ (p) min{p − q : q ∈ bndry(Q)}.
Critical Points and Directions of Descent Note that the value of smQ at p is the radius of the largest
A critical point of f : Rd → R is a point x ∈ Rd such that disk with center p contained in the polygon. Moreover, the
0 ∈ ∂ f (x). According to this definition, the maximizers and function smQ is locally Lipschitz on Q. To show this,
minimizers of a locally Lipschitz function are critical rewrite smQ as
points. As an example, x = 0 is the unique minimizer of
f (x) = |x|, and, indeed, we see that 0 ∈ ∂ f (0) in (45).
smQ (p) min{dist(p, e) : e is an edge of Q},
If a function f is continuously differentiable, then the gra-
dient ∇ f provides the direction of maximum ascent of f
(respectively, −∇ f provides the direction of maximum where dist(p, e) denotes the Euclidean distance from the
descent of f ). When we consider a locally Lipschitz function, point p to the edge e. Indeed, the function smQ is concave
however, the question of choosing the directions of descent on Q.
among all of the available directions in the generalized gra- Let us consider the generalized gradient vector field
dient arises. Without loss of generality, we restrict our dis- corresponding to smQ , where the definition of smQ is
cussion to directions of descent, since a direction of descent extended outside Q by setting smQ (p) = − min{p − q2 :
of − f corresponds to a direction of ascent of f , while f is q ∈ bndry (Q)} for p ∈
/ Q. Applying Proposition 7-iii), we
locally Lipschitz if and only if − f is locally Lipschitz. deduce that −smQ is regular on Q and its generalized
Let Ln : B(Rd) → B(Rd) be the set-valued map that gradient at p ∈ Q is
associates to each subset S of Rd the set of least-norm ele-
ments of its closure S. If the set S is convex and closed, ∂smQ (p) = co{ne : e edge of Q
then the set Ln(S) is a singleton, which consists of the such that smQ (p) = dist(p, e)},
orthogonal projection of zero onto S. For a locally Lipschitz
function f , consider the generalized gradient vector field where ne denotes the unit normal to the edge e pointing
Ln(∂ f ) : Rd → Rd defined by toward the interior of Q. Therefore, at points p in Q for
which there exists a unique edge e of Q nearest to p, the
function smQ is differentiable, and its generalized gradient
x → Ln(∂ f )(x) Ln(∂ f (x)).
vector field is given by Ln(smQ )(p) = ne . Note that this
vector field corresponds to the move-away-from-nearest-
It turns out that −Ln(∂ f )(x) is a direction of descent of f at neighbor interaction law for one agent moving in the poly-
x ∈ Rd . More precisely [9], if 0 ∈ / ∂ f (x), then there exists gon introduced in Example 4!
T > 0 such that At points p of Q where various edges {e1 , . . . , em } are at
the same minimum distance to p, the function smQ is not
t differentiable, and its generalized gradient vector field is
f (x − tLn(∂ f )(x)) ≤ f (x) − Ln(∂ f )(x)22 < f (x),
2 given by the least-norm element in co{ne1 , . . . , nem }. If p is
0 < t < T, (47) not a critical point, zero does not belong to co{ne1 , . . . , nem },
and the least-norm element points in the direction of the
that is, by taking a small step in the direction −Ln(∂ f )(x), bisector line between two of the edges in {e1 , . . . , em }.
the function f is guaranteed to decrease by an amount that Figure 9 shows a plot of the generalized gradient vector
scales linearly with the stepsize. field of smQ on the square Q = [−1, 1]2 . Note the similarity
x2
0
Given a locally Lipschitz function f : Rd → R , the non-
smooth analog of the classical gradient descent flow of a
differentiable function is defined by
−.5
ẋ(t) = −Ln(∂ f )(x(t)). (48)
epi( f ) = {(x, μ) ∈ Rd × R : f (x) ≤ μ} ⊂ Rd+1 . The function y ∈ Rd such that x is the nearest point in S to x + λy, for
f is lower semicontinuous if and only if its epigraph is λ > 0 sufficiently small. Then, ζ ∈ ∂P f (x) if and only if
closed. The function f : Rd → R is upper semicontinuous at
x ∈ Rd if − f is lower semicontinuous at x. Note that f is (ζ, −1) ∈ Nepi(f) (x, f (x)). (52)
continuous at x if and only if f is both upper semicontinu-
ous and lower semicontinuous at x. Figure 10(b) illustrates this geometric interpretation.
For a lower semicontinuous function f : Rd → R , the
vector ζ ∈ Rd is a proximal subgradient of f at x ∈ Rd if there Example 17: Proximal Subdifferentials of the
exist σ, δ ∈ (0, ∞) such that, for all y ∈ B(x, δ), Absolute Value Function and Its Negative
Consider the locally Lipschitz functions f, g : R → R ,
f (y) ≥ f (x) + ζ(y − x) − σ 2 ||y − x||22 . (51) f (x) = |x| and g(x) = −|x|. Using the geometric interpreta-
tion of (51), it can be seen that
The set ∂P f (x) of all proximal subgradients of f at x is the
proximal subdifferential of f at x. The proximal subdifferen- ⎧
⎨ {1}, x < 0,
tial at x, which may be empty, is convex but not necessari- ∂P f (x) = [−1, 1], x = 0, (53)
ly open, closed, or bounded. ⎩
{−1}, x > 0,
Geometrically, the definition of a proximal subgradient ⎧
can be interpreted as follows. Equation (51) is equivalent to ⎨ {−1}, x < 0,
∂P g(x) = ∅, x = 0, (54)
saying that, in a neighborhood of x, the function y → f (y) ⎩
{1}, x > 0.
majorizes the quadratic function y → f (x)+ ζ(y − x)
−σ 2 y − x22 . In other words, there exists a parabola that
locally fits under the epigraph of f at (x, f (x)). This geo- Note that (53)–(54) is an example of the fact that ∂P (− f )
metric interpretation is useful for explicitly computing the and −∂P f are not necessarily equal. The generalized gradi-
proximal subdifferential; see Figure 10(a) for an illustra- ent of g in (46) is different from the proximal subdifferen-
tion. An additional geometric characterization of the proxi- tial of g in (54).
mal subdifferential of f can be given in terms of normal Unlike the case of the generalized gradient, the prox-
cones and the epigraph of f . Given a closed set S ⊆ Rd and imal subdifferential may not coincide with ∇ f (x) when
x ∈ S, the proximal normal cone NS (x) to S is the set of all f is continuously differentiable. The function f : R → R,
where f (x) = −|x|3/2 , is continu-
ously differentiable with
∇ f (0) = 0, but ∂P f (0) = ∅ . In
fact, a continuously differen-
tiable function on R with an
empty proximal subdifferential
1
almost everywhere is provided
in [48]. However, the density
–1 1
theorem (cf. [24, thrm. 3.1]) states
that the proximal subdifferential
–1 of a lower semicontinuous func-
tion is nonempty on a dense set
of its domain of definition,
(a) (b) although a dense set can have
zero Lebesgue measure.
FIGURE 10 Geometric interpretations of the proximal subdifferential of the function x → |x| at On the other hand, the proxi-
x = 0 computed in Example 17. The epigraph of the function is shaded, while the proximal nor-
mal subdifferential can be more
mal cone to the epigraph at zero is striped. In (a), according to (51) each proximal subgradient
corresponds to a direction tangent to a parabola that fits under the epigraph of the function. In useful than the generalized gradi-
(b), according to (52), each proximal subgradient can be uniquely associated with an element of ent in some situations, as the fol-
the proximal normal cone to the epigraph of the function. lowing example shows.
and proximal subdifferential, the relevant notions are to Filippov solutions by taking F = F[X], and to control
upper semicontinuous function, concave function, and systems by taking F = G[X].
proximal super differential, respectively [24]. Before proceeding with our exposition, we recall that
solutions of a discontinuous system are not necessarily
Gradient Differential Inclusion of a Convex Function unique. Therefore, when considering a property such as
It is not always possible to associate a nonsmooth gradient Lyapunov stability, we must specify whether attention is
flow with a lower semicontinuous function because the being paid to a particular solution starting from an initial
proximal subdifferential might be empty almost every- condition (“weak”) or to all the solutions starting from an
where. However, following Proposition 9-ii), we can asso- initial condition (“strong”). As an example, the set M ⊆ Rd
ciate a nonsmooth gradient flow with each convex is weakly invariant for (63) if, for each x0 ∈ M, M contains at
function, as we briefly discuss next [49]. least one maximal solution of (63) with initial condition x0 .
Let f : Rd → R be convex, and consider the gradient Similarly, M ⊆ Rd is strongly invariant for (63) if, for each
differential inclusion x0 ∈ M, M contains every maximal solution of (63) with
initial condition x0 . Rather than reintroducing weak and
ẋ(t) ∈ −∂P f (x(t)). (62) strong versions of standard stability notions, we rely on
the guidance provided above and the context for the spe-
Using the properties of the proximal subdifferential given cific meaning in each case. Detailed definitions are given in
by Proposition 9-ii), the existence of a Caratheodory solu- [14] and [18].
tion of (62) starting from every initial condition is guaran- The following discussion requires the notion of a limit
teed by Proposition S2. Moreover, the uniqueness of point of a solution of a differential inclusion. The point
Caratheodory solutions can be established as follows. Let x∗ ∈ Rd is a limit point of a solution t → x(t) of (63) if there
x, y ∈ Rd , and take ζ1 ∈ −∂P f (x) and ζ2 ∈ −∂P f (y). Using exists a sequence {tn }n∈N such that x(tn ) → x∗ as n → ∞.
Proposition 9-i), we have We denote by (x) the set of limit points of t → x(t).
Under the hypotheses of Proposition S2, (x) is a weakly
f (y) ≥ f (x) − ζ1 (y − x), f (x) ≥ f (y) − ζ2 (x − y). invariant set. Moreover, if the solution t → x(t) lies in a
bounded, open, and connected set, then (x) is nonempty,
From here, we deduce −ζ1 (y − x) ≤ f (y) − f (x) ≤ bounded, connected, and x(t) → (x) as t → ∞; see [18].
−ζ2 (y − x), and therefore (ζ2 − ζ1 )(y − x) ≤ 0, which, in
particular, implies that the set-valued map x → −∂P f (x) Stability Analysis by Means of the Generalized
satisfies the one-sided Lipschitz condition (S3). Gradient of a Nonsmooth Lyapunov Function
Proposition S3 then guarantees that there exists a unique In this section, we discuss nonsmooth stability analysis
Caratheodory solution of (62) starting from every initial results that use locally Lipschitz functions and generalized
condition. gradients. We present results taken from various sources
in the literature. This discussion is not intended to be a
NONSMOOTH STABILITY ANALYSIS comprehensive account of such a vast topic but rather
In this section, we study the stability of discontinuous serves as a motivation for further exploration. The books
dynamical systems. Unless explicitly mentioned otherwise, [14], [18] and journal articles [51]–[54] provide additional
the stability notions employed here correspond to the usual information.
ones for differential equations; see, for instance, [50]. The pre-
sentation focuses on the time-invariant differential inclusion Lie Derivative and Monotonicity
A common theme in stability analysis is establishing the
ẋ(t) ∈ F(x(t)), (63) monotonic evolution of a candidate Lyapunov function
along the trajectories of the system. Mathematically, the
where F : Rd → B(Rd) . Throughout this section, we evolution of a function along trajectories is captured by the
assume that the set-valued map F satisfies the hypotheses notion of Lie derivative. Our first task is to generalize this
of Proposition S2, so that the existence of solutions of (63) notion to the setting of differential inclusions following
is guaranteed. The scenario (63) has a direct application to [53]; see also [51] and [52].
discontinuous differential equations and control systems. Given a locally Lipschitz function f : Rd → R and a set-
For instance, the results presented here apply to valued map F : Rd→ B(Rd), the set-valued Lie derivative
Caratheodory solutions by taking a singleton-valued map, L̃F f : Rd → B(R) of f with respect to F at x is defined as
{(sign(x1 ), x2 )}, x1 =
0,
L̃F f (x) = {a ∈ R : there exists v ∈ F(x) such that ∂ f (x1 , x2 ) =
[−1, 1] × {x2 }, x1 = 0.
ζ T v = a for all ζ ∈ ∂ f (x)}. (64)
0
weakly invariant set M contained in
0 (67)
FIGURE 12 From left to right, evolution of the nonsmooth gradient flow of the function −sm Q in a convex polygon. At each snapshot, the value
of sm Q is the radius of the largest disk (plotted in gray) contained in the polygon with center at the current location. The flow converges in
finite time to the incenter set, which, for this polygon, is a singleton whose only element is the center of the disk in the rightmost snapshot.
FIGURE 13 The move-away-from-nearest-neighbor interaction law for solving a sphere-packing problem within the polygon Q. (a) The initial
configuration, (b) the evolution, and (c) the final configuration of a Filippov solution. In (c), the minimum radius of the shaded spheres corre-
sponds to the value of the locally Lipschitz function HSP , defined in (69). The Filippov solutions of the move-away-from-nearest-neighbor
dynamical system monotonically increase the value of HSP . In (c), every node is at equilibrium since the infinitesimal motion of a node in any
direction would place it closer to at least another node. This discontinuous dynamical system [44] is an example of how simple local interac-
tions can achieve a global objective.
Lie Derivative and Monotonicity the role played by the set-valued Lie derivative L̃F f for a
Let D ⊆ Rd be an open and connected set. A lower semi- locally Lipschitz function. These objects allow us to study
continuous function f : Rd → R is weakly nonincreasing on how f evolves along the solutions of a differential inclu-
D for a set-valued map F : Rd → B(Rd) if, for all y ∈ D, sion without having to obtain the solutions in closed form.
there exists a solution x : [0, t1 ] → Rd of the differential Specifically, we have the following result [24]. In this and
inclusion (63) starting at y and contained in D that satisfies in forthcoming statements, it is convenient to adopt the
convention sup ∅ = −∞.
f (x(t)) ≤ f (x(0))
Proposition 13
= f (y) for all t ∈ [0, t1 ].
Let F : Rd → B(Rd) be a set-valued map satisfying the
hypotheses of Proposition S2, and consider the associated
If, in addition, f is continuous, then being weakly non- differential inclusion (63). Let f : Rd → R be a lower semi-
increasing is equivalent to the property of having a solu- continuous function, and let D ⊆ Rd be open. Then, the
tion starting at y such that t → f (x(t)) is monotonically following statements hold:
nonincreasing on [0, t1 ]. i) The function f is weakly nonincreasing on D if and
Similarly, a lower semicontinuous function f : Rd → R only if
is strongly nonincreasing on D for a set-valued map
F : Rd → B(Rd) if, for all y ∈ D , all solutions sup LF f (y) ≤ 0, for all y ∈ D.
x : [0, t1 ] → Rd of the differential inclusion (63) starting at
y and contained in D satisfy ii) If, in addition, either F is locally Lipschitz on D, or F
is continuous on D and f is locally Lipschitz on D,
then f is strongly nonincreasing on D if and only if
f (x(t)) ≤ f (x(0))
= f (y) for all t ∈ [0, t1 ].
sup LF f (y) ≤ 0, for all y ∈ D.
ferentiable on the open right and left half-planes, together ⎨ −σ ((x1 , x2 )2 ) 21 22 , x1 = 0,
x1 +x2 +|x1 |
LFσ f (x1 , x2 ) =
with the geometric interpretation of proximal subgradi- ⎩
ents, we obtain −∞, x1 = 0,
(75)
⎧ (x2 +x2 )
3/2
⎨ σ ((x1 , x2 )2 ) 21 22 , x1 = 0,
∂P f (x⎧
1 , x2 )
⎧ ⎪⎛ ⎞⎫ LFσ f (x1 , x2 ) = x1 +x2 +|x1 |
⎨ ⎪
⎬ ⎩
⎪
⎪ ⎜ x1 +x2 −2x1 x1 +x2 2 1
2 2 2 2 x 2x + x 2
+x2
2 ⎟ −∞, x1 = 0.
⎪
⎪ ⎝− 2 2 ,
1
⎠ , x1 > 0,
⎪
⎪ ⎪ 2
⎪
⎪
⎪ ⎩ x1 +x2 +x1 x21 +x22 x1 + x21 +x22 ⎭ (76)
⎪
⎪
⎪
⎨
= ∅, x1 = 0, Therefore, sup LFσ f (x1 , x2 ) ≤ 0 for all (x1 , x2 ) ∈ R2 . Now
⎪
⎪
⎪
⎪ ⎧⎛ ⎞⎫ using Proposition 13-i), we deduce that f is weakly
⎪
⎪ ⎪ ⎪
⎪ ⎨
⎪ ⎜ x1 +x2 +2x1 x1 +x2 x −2x + x2
+x 2
2 ⎟
⎬ nonincreasing on R2 . Since f is continuous, this fact is
⎪
⎪
2 2 2 2
,
2 1 1
⎪ ⎝ 2 2
⎩ 2 ⎠ , x1 < 0. equivalent to saying that there exists a control input u
⎪
⎩ x1 +x2 −x1 x1 +x2
2 2
x1 − x21 +x22
⎪
⎭
such that the solution t → x(t) of the dynamical system
x2
2
2
2
1
1 1
0
x2
0 x1
x2
−1 −1
−1
−2
−2 −2
−2 −1 0 1 2
−2 −1 0 1 2 −3 x1
x1 (c)
(a) (b)
the input vector field (x1 , x 2 ) → (x1 − x 2 , 2x1 x 2 ), (b) its integral curves, and (c) the con-
2 2
FIGURE 14 Cart on a circle. (a) The phase portrait of
tour plot of the function 0 = (x1 , x 2 ) → (x1 + x 2 /( x1 + x 2 + |x1 |)), (0, 0) → 0. The origin cannot be asymptotically stabilized by means of
2 2 2 2
continuous feedback in the cart dynamics (72)–(73). However, the origin can be asymptotically stabilized by means of discontinuous feedback
when solutions are understood in the sample-and-hold sense.
resulting from substituting u in (72)–(73) has the prop- f, g : Rd → R are a Lyapunov pair for an equilibrium
erty that t → f (x(t)) is monotonically nonincreasing. xe ∈ Rd if they satisfy f (x), g(x) ≥ 0 for all x ∈ Rd , and
g(x) = 0 if and only if x = xe ; f is radially unbounded,
Stability Results and, moreover,
The results presented in the previous section establishing
the monotonic behavior of a lower semicontinuous func- sup LF f (x) ≤ −g(x) for all x ∈ Rd.
tion allow us to provide tools for stability analysis. We pre-
sent here an exposition parallel to the discussion for the If an equilibrium xe of (63) admits a Lyapunov pair, then there
locally Lipschitz function and generalized gradient case. exists at least one solution starting from every initial condition
We start by presenting a result on Lyapunov stability that that asymptotically converges to the equilibrium; see [24].
follows from the exposition in [24].
Example 20 Revisited: Asymptotic Stability
Theorem 3 of the Origin in the Cart Example
Let F : Rd → B(Rd) be a set-valued map satisfying the As an application of the above discussion and the version
hypotheses of Proposition S2. Let xe be an equilibrium of of Theorem 3 for weak stability, consider the cart on a circle
the differential inclusion (63), and let D ⊆ Rd be a domain introduced in Example 20. Setting xe = (0, 0) and D = R2 ,
with xe ∈ D. Let f : Rd → R and assume that the following and taking into account the computation (75) of the lower
conditions hold: set-valued Lie derivative, we conclude that (0, 0) is a glob-
i) F is continuous on D and f is locally Lipschitz on D, ally weakly asymptotically stable equilibrium.
or F is locally Lipschitz on D and f is lower semicon- We now turn our attention to the extension of the
tinuous on D, and f is continuous at xe . invariance principle for differential inclusions using the
ii) f (xe ) = 0, and f (x) > 0 for x ∈ D\{xe }. proximal subdifferential of a lower semicontinuous func-
iii) sup LF f (x) ≤ 0 for all x ∈ D. tion. The following result can be derived from the exposi-
Then, xe is a strongly stable equilibrium of (63). In tion in [24].
addition, if iii) above is replaced by
iii)’ sup LF f (x) ≤ 0 for all x ∈ D\{xe }, Theorem 4
then xe is a strongly asymptotically stable equilibrium Let F : Rd → B(Rd) be a set-valued map satisfying the
of (63). hypotheses of Proposition S2, and let f : Rd → R. Assume
A similar result can be stated for weakly stable that either F is continuous and f is locally Lipschitz, or F
equilibria substituting i) by “i)’ f is continuous on D,” is locally Lipschitz and f is continuous. Let S ⊂ Rd be
and the upper set-valued Lie derivative by the lower compact and strongly invariant for (63), and assume that
set-valued Lie derivative in iii) and iii)’. Note that, if sup LF f (y) ≤ 0 for all y ∈ S. Then, every solution
the differential inclusion (63) has a unique solution x : [0, ∞) → Rd of (63) starting at S converges to the
starting from every initial condition, then the notions largest weakly invariant set M contained in
of strong and weak stability coincide, and it is suffi-
cient to satisfy the simpler requirements i)’ and iii)’ for S ∩ {y ∈ Rd : 0 ∈ LF f (y)}.
weak stability.
Similarly to the case of a continuous differential equa- Moreover, if the set M consists of a finite number of points,
tion, global asymptotic stability can be established by then the limit of each solution starting in S exists and is an
requiring the Lyapunov function f to be continuous and element of M.
radially unbounded. The equivalence between global Next, we apply Theorem 4 to gradient differential
strong asymptotic stability and the existence of infinitely inclusions.
differentiable Lyapunov functions is discussed in [62].
This type of global result is commonly invoked when Stability of Gradient Differential
dealing with the stabilization of control systems by refer- Inclusions for Convex Functions
ring to control Lyapunov functions [25] or Lyapunov Consider the gradient differential inclusion (62) associat-
pairs [24]. Two lower semicontinuous functions ed with the convex function f : Rd → R. We study here