Skip to main content
Log in

Finite-time consensus protocols for multi-dimensional multi-agent systems

  • Research article
  • Published:
Control Theory and Technology Aims and scope Submit manuscript

Abstract

A finite-time consensus protocol is proposed for multi-dimensional multi-agent systems, using direction-preserving signum controls. Filippov solutions and nonsmooth analysis techniques are adopted to handle discontinuities. Sufficient and necessary conditions are provided to guarantee infinite-time convergence and boundedness of the solutions. It turns out that the number of agents which have continuous control law plays an essential role in finite-time convergence. In addition, it is shown that the unit balls introduced by \(\ell _p\) norms, where \(p\in [1,\infty ]\), are invariant for the closed loop.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. The definition of a regular function can be found in [26]

References

  1. Jadbabaie, A., Lin, J., & Morse, A.S. (2003). Coordination of groups of mobile autonomous agents using nearest neighbor rules. IEEE Transactions on Automatic Control, 48(6), 988–1001.

    Article  MathSciNet  Google Scholar 

  2. Olfati-Saber, R., Fax, J. A., & Murray, R. M. (2007). Consensus and cooperation in networked multi-agent systems. Proceedings of the IEEE, 95(1), 215–233.

    Article  Google Scholar 

  3. Ren, W., Beard, R.W., & Atkins, E.M. (2005). A survey of consensus problems in multi-agent coordination. In Proceedings of the American Control Conference (pp. 1859–1864). Portland, OR.

  4. Bhat, S.P. & Bernstein, D. S. (2000). Finite-time stability of continuous autonomous systems. SIAM Journal on Control and Optimization, 38(3), 751–766.

    Article  MathSciNet  Google Scholar 

  5. Cao, Y., & Ren, W. (2014). Finite-time consensus for multi-agent networks with unknown inherent nonlinear dynamics. Automatica, 50(10), 2648–2656.

    Article  MathSciNet  Google Scholar 

  6. Hui, Q., Haddad, W. M., & Bhat, S. P. (2008). Finite-time semistability and consensus for nonlinear dynamical networks. IEEE Transactions on Automatic Control, 53(8), 1887–1900.

    Article  MathSciNet  Google Scholar 

  7. Wang, L., & Xiao, F. (2010). Finite-time consensus problems for networks of dynamic agents. IEEE Transactions on Automatic Control, 55(4), 950–955.

    Article  MathSciNet  Google Scholar 

  8. Xiao, F., Wang, L., Chen, J., & Gao, Y. (2009). Finite-time formation control for multi-agent systems. Automatica, 45(11), 2605–2611.

    Article  MathSciNet  Google Scholar 

  9. Harl, N., & Balakrishnan, S. N. (2012). Impact time and angle guidance with sliding mode control. IEEE Transactions on Control Systems Technology, 20(6), 1436–1449.

    Article  Google Scholar 

  10. Chen, F., Cao, Y., & Ren., W. (2012). Distributed average tracking of multiple time-varying reference signals with bounded derivatives. IEEE Transactions on Automatic Control, 57(12), 3169–3174.

    Article  MathSciNet  Google Scholar 

  11. Chen, G., Lewis, F. L., & Xie, L. (2011). Finite-time distributed consensus via binary control protocols. Automatica, 47(9), 1962–1968.

    Article  MathSciNet  Google Scholar 

  12. Cortés, J. (2006). Finite-time convergent gradient flows with applications to network consensus. Automatica, 42(11), 1993–2000.

    Article  MathSciNet  Google Scholar 

  13. Franceschelli, M., Giua, A., & Pisano, A. (2017). Finite-time consensus on the median value with robustness properties. IEEE Transactions on Automatic Control, 62(4), 1652–1667.

    Article  MathSciNet  Google Scholar 

  14. Hui, Q., Haddad, W. M., & Bhat, S. P. (2010). Finite-time semistability, filippov systems, and consensus protocols for nonlinear dynamical networks with switching topologies. Nonlinear Analysis: Hybrid Systems, 4(3), 557–573.

    MathSciNet  MATH  Google Scholar 

  15. Li, C., & Qu, Z. (2014). Distributed finite-time consensus of nonlinear systems under switching topologies. Automatica, 50(6), 1626–1631.

    Article  MathSciNet  Google Scholar 

  16. Liu, X., Lam, J., Yu, W., & Chen, G. (2016). Finite-time consensus of multiagent systems with a switching protocol. IEEE Transactions on Neural Networks and Learning Systems, 27(4), 853–862.

    Article  MathSciNet  Google Scholar 

  17. Wei, J., Everts, A. R.F., Camlibel, M. K., & van der Schaft, A. J. (2017). Consensus dynamics with arbitrary sign-preserving nonlinearities. Automatica, 83, 226–233.

    Article  MathSciNet  Google Scholar 

  18. Wei, J., Zhang, S., Adaldo, A., Thunberg, J., Hu, X., & Johansson, K.H. (2018). Finite-time attitude synchronization with distributed discontinuous protocols. IEEE Transactions on Automatic Control, 63(10), 3608–3615.

  19. Thunberg, J., Song, W., Montijano, E., Hong, Y., Hu, X. (2014). Distributed attitude synchronization control of multi-agent systems with switching topologies. Automatica, 50(3), 832–840.

    Article  MathSciNet  Google Scholar 

  20. Shames, I., Dasgupta, S., Fidan, B., & Anderson, B. D. O. (2012). Circumnavigation using distance measurements under slow drift. IEEE Transactions on Automatic Control, 57(4), 889–903.

    Article  MathSciNet  Google Scholar 

  21. Biggs, N. (1993). Algebraic Graph Theory. 2nd ed. Cambridge: Cambridge University Press.

  22. Bollobas, B. (1998). Modern Graph Theory. Graduate Texts in Mathematics, Vol. 184. New York: Springer.

  23. Filippov, A. F. (1988). Differential Equations with Discontinuous Righthand Sides: Control Systems. Arscott, F. M. (ed.). Mathematics and its Applications. Netherlands: Springer.

  24. van der Schaft, A. J. (2010). Characterization and partial synthesis of the behavior of resistive circuits at their terminals. Systems & Control Letters, 59(7), 423–428.

    Article  MathSciNet  Google Scholar 

  25. Cortés, J. (2008). Discontinuous dynamical systems. IEEE Control Systems, 28(3), 36–73.

    Article  MathSciNet  Google Scholar 

  26. Clarke, F.H. (1990). Optimization and Nonsmooth Analysis. Classics in Applied Mathematics. Society for Industrial and Applied Mathematics.

  27. Hale, J. K. (2009). Ordinary Differential Equations. Dover Books on Mathematics Series. Dover Publications.

  28. Khalil, H. K. (2002). Nonlinear Systems. Upper Saddle River: Prentice Hall.

  29. Bacciotti, A., & Ceragioli, F. (1999). Stability and stabilization of discontinuous systems and nonsmooth Lyapunov functions. ESAIM: Control, Optimisation and Calculus of Variations, 4, 361–376.

  30. Paden, B., & Sastry, S. (1987). A calculus for computing Filippov’s differential inclusion with application to the variable structure control of robot manipulators. IEEE Transactions on Circuits and Systems, 34(1), 73–82.

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jieqiang Wei.

Appendix

Appendix

Proof of Lemma 2

We divide the proof into two parts, discussing the cases \(p\in [1,\infty )\) and \(p=\infty\) separately.

  1. 1)

    Let \(p\in [1,\infty )\). We introduce a Lyapunov function candidate

    $$\begin{aligned} V(x) = \max _{i\in {\mathcal {I}}} \dfrac{1}{p}\Vert x_i\Vert _p^p \end{aligned}$$
    (A1)

    and note that \(V(x)\leqslant \dfrac{1}{p}C^p\) implies that \(x\in {\mathcal {S}}_p(C)\). Since the function \((\cdot )^p\) is convex on \(\mathbb {R}_{\geqslant 0}\), it can be observed that V is convex and, hence, regular. In the remainder of the proof, we will show that V(x(t)) is nonincreasing along all Filippov solutions of (14), implying strong invariance of the set \({\mathcal {S}}_p(C)\) for any \(C>0\).

Let \(\alpha (x)\) denote the set of indices that achieve the maximum in (45) as

$$\begin{aligned} \alpha (x) = \{ i\in {\mathcal {I}}\,|\,\Vert x_i\Vert _p^p = pV(x) \}. \end{aligned}$$
(A2)

Then, the generalized gradient of V in (45) is given by

$$\begin{aligned} \partial V(x)=\mathrm {co} \{ e_i\otimes \psi (x_i)\,|\,i\in \alpha (x) \} \end{aligned}$$
(A3)

where

$$\begin{aligned} \psi (x_i) = \left[ \begin{array}{c} |x_{i,1}|^{p-1}{\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,1}) \\ \vdots \\ |x_{i,k}|^{p-1}{\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,k})\end{array}\right] \end{aligned}$$
(A4)

and \(x_i = [x_{i,1} ~~\cdots ~~ x_{i,k}]^{\mathrm{T}}\in \mathbb {R}^k\).

Next, let \(\varPsi\) be defined as

$$\begin{aligned} \varPsi = \Big\{ t\geqslant 0\,\big|\,{\dot{x}}(t) \text { and } \dfrac{\mathrm{d}}{{\mathrm{d}}t}V(x(t)) \text { exist} \Big\} . \end{aligned}$$
(A5)

Since x is absolutely continuous (by definition of Filippov solutions) and V is locally Lipschitz, by Lemma 1 in [29] it follows that \(\varPsi =\mathbb {R}_{\geqslant 0}{\setminus }{\bar{\varPsi }}\) for a set \({\bar{\varPsi }}\) of measure zero and

$$\begin{aligned} \dfrac{\mathrm{d}}{{\mathrm{d}}t}V(x(t))\in {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t)) \end{aligned}$$
(A6)

for all \(t\in \varPsi\), such that the set \({\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t))\) is nonempty for all \(t\in \varPsi\). For \(t\in {\bar{\varPsi }}\), we have that \({\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t))\) is empty, and hence \(\max {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t)) = -\infty < 0\) by definition. Therefore, we only consider \(t\in \varPsi\) in the rest of the proof.

Next, we will consider the cases \(x\in {\mathcal {C}}\) and \(x\notin {\mathcal {C}}\) separately.

First, for \(x\in {\mathcal {C}}\), it can be observed that \(\alpha (x) = {\mathcal {I}}\). Then, the following two cases can be distinguished.

  1. i)

    \(|{\mathcal {I}}| \geqslant 2\) and \(|{\mathcal {I}}_c|\geqslant 1\). As there is at least one agent with continuous vector field, there exists \(i\in {\mathcal {I}}\) such that \(f_i\) is locally Lipschitz and direction preserving. Then, by definition of the Filippov set-valued map, it follows that \(\nu _i = 0\) for all \(\nu \in {\mathcal {F}}[h](x)\) (recall \(x\in {\mathcal {C}}\)). As \({\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t))\) is nonempty (by considering \(t\in \varPsi\)), there exists \(a\in {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t))\) such that \(a = \zeta ^{\mathrm{T}}\nu\) for all \(\zeta \in \partial V(x(t))\), see definition (6). Choosing \(\zeta = e_i\otimes \psi (x_i(t))\), it follows that \(a = (e_i\otimes \psi (x_i(t)))^{\mathrm{T}}\nu = 0\), which implies that \(\max {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t)) = 0 \leqslant 0\), i.e., V(x) is nonincreasing for any \(x\in {\mathcal {C}}\).

  2. ii)

    \(|{\mathcal {I}}| = 2\) and \(|{\mathcal {I}}_c| = 0\). In this case, system (13) can be written as

    $$\begin{aligned} \left\{ \begin{aligned} {\dot{x}}_1&= \dfrac{x_2-x_1}{\Vert x_2-x_1\Vert _p}, \\ {\dot{x}}_2&= \dfrac{x_1-x_2}{\Vert x_1-x_2\Vert _p}. \end{aligned}\right. \end{aligned}$$
    (A7)

    Then, by using definition (3), it can be shown that, for \(x_1 = x_2\) (i.e., \(x\in {\mathcal {C}}\)), any element \(\nu\) in the Filippov set-valued map of (51) satisfies \(\nu _1 = -\nu _2\). Stated differently, the following implication holds with \(\nu = [\nu _1^{\mathrm{T}}~~\nu _2^{\mathrm{T}}]^{\mathrm{T}}\):

    $$\begin{aligned} \nu \in {\mathcal {F}}[h](x),\; x\in {\mathcal {C}} \;\Rightarrow \; \nu _1 = -\nu _2. \end{aligned}$$
    (A8)

    Next, by recalling that \(\alpha (x) = {\mathcal {I}}\) (see (46)), it follows from (47) that

    $$\begin{aligned} \partial V(x) = \mathrm {co}\{e_1\otimes \psi (x_1(t)), e_2\otimes \psi (x_2(t))\} \end{aligned}$$
    (A9)

    with \(x_1 = x_2\). Now, following a similar reasoning as in item (i) on the basis of the definition of the set-valued Lie derivative in (6), it can be concluded that \(a = \zeta ^{\mathrm{T}}\nu\) is necessarily 0, such that \(\max {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t)) = 0 \leqslant 0\) for all \(x\in {\mathcal {C}}\).

Second, the case \(x\notin {\mathcal {C}}\) is considered. For this case, Theorem 1 in [30] is applied to obtain

$$\begin{aligned} {\mathcal {F}}[h](x) \subset \times _{j=1}^N {\mathcal {F}}[f_i](-{\bar{L}}_ix) =: \bar{{\mathcal {F}}}(x), \end{aligned}$$
(A10)

after which it follows from the definition of the set-valued Lie derivative (6) that

$$\begin{aligned} {\mathcal {L}}_{{\mathcal {F}}}V(x) \subset {\mathcal {L}}_{\bar{{\mathcal {F}}}}V(x). \end{aligned}$$
(A11)

Therefore, in the remainder of the proof for the case \(x\notin {\mathcal {C}}\), we will show that \(\max {\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x(t)) \leqslant 0\), which implies the desired result by (55). As before, it is sufficient to consider the set \(\varPsi\) such that \({\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x(t))\) is nonempty for all \(t\in \varPsi\).

Now, take an index \(i\in \alpha (x)\) such that \({\bar{L}}_ix \ne {\mathbf {0}}\). Note that such i indeed exists. Namely, assume in order to establish a contradiction that \({\bar{L}}_ix = {\mathbf {0}}\) for all \(i\in \alpha (x)\). If \(\alpha (x)={\mathcal {I}}\), then there exists \(\ell \in \{1,\ldots ,k\}\) such that

$$\begin{aligned} \beta (\ell ):=\arg \max _{j\in {\mathcal {I}}}x_{j,\ell } \subsetneq {\mathcal {I}}, \end{aligned}$$
(A12)

i.e., there exists a state component \(\ell\) that does not have the same value for all agents. Otherwise, \(x\in {\mathcal {C}}\), which is a contradiction. Then, for any \(i\in \beta (\ell )\) with \(j\in N_i{\setminus } \beta (\ell )\), we have \({\bar{L}}_ix \ne {\mathbf {0}}\), where \(N_i\) is the set of neighbors of agent i. If \(\alpha (x)\subsetneq {\mathcal {I}}\) and \({\bar{L}}_ix = {\mathbf {0}}\) for all \(i\in \alpha (x)\), then for any \(i\in \alpha (x)\) with \(j\in N_i{\setminus }\alpha (x)\), we have

$$\begin{aligned} 0= \psi^{\mathrm{T}} (x_i){\bar{L}}_ix \end{aligned}$$
(A13)
$$\begin{aligned}&\quad= \textstyle \sum \limits _{j\in N_i} (\Vert x_i\Vert _p^p - \textstyle \sum \limits _{\ell =1}^{k}|x_{i,\ell }|^{p-1}{\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })x_{j,\ell }) \nonumber \\& \quad\geqslant \textstyle \sum \limits _{j\in N_i} (\Vert x_i\Vert _p^p - \textstyle \sum \limits _{\ell =1}^{k}|x_{i,\ell }|^{p-1} |x_{j,\ell }|) \nonumber \\&\quad \geqslant \textstyle \sum \limits _{j\in N_i} (\Vert x_i\Vert _p^p - \Vert x_j\Vert _p \Vert x_i\Vert _p^{p-1} )\nonumber \\&\quad > 0 ,\end{aligned}$$
(A14)

where inequality (58) is based on Hölder’s inequality, and the last inequality is implied by \(\Vert x_i\Vert >\Vert x_j\Vert\) for any \(j\in N_i{\setminus }\alpha (x)\). This is a contradiction.

For the index \(i\in \alpha (x)\) satisfying \({\bar{L}}_ix\ne {\mathbf {0}}\), it follows from Assumption 1 that there exists \(\gamma >0\) such that

$$\begin{aligned} {\mathcal {F}}[f_i](-{\bar{L}}_ix) = \{-\gamma {\bar{L}}_ix \}, \end{aligned}$$
(A15)

i.e., for any \(\nu \in \bar{{\mathcal {F}}}(x)\) it holds that \(\nu _i = -\gamma {\bar{L}}_ix\). Note that this is a result of the direction-preserving property of either the direction-preserving signum (for a nonzero argument, then \(\gamma = \dfrac{1}{\Vert {\bar{L}}_ix\Vert _p})\) or the Lipschitz continuous function (by Assumption 1). Then, choosing \(\zeta \in \partial V(x)\) as \(\zeta = e_i\otimes \psi (x_i)\) (recall that \(i\in \alpha (x)\)), it follows from (6) that

$$\begin{aligned} {\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x) = \{-\gamma \psi ^{\mathrm{T}}(x_i){\bar{L}}_ix\}. \end{aligned}$$
(A16)

Next, by observing (58), we have

$$\begin{aligned} \psi ^{\mathrm{T}}(x_i){\bar{L}}_ix \geqslant 0, \end{aligned}$$
(A17)

which implies \({\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x)\subset \mathbb {R}_{\leqslant 0}\).

Summarizing the results of the two cases leads to the condition

$$\begin{aligned} \max {\mathcal {L}}_{{\mathcal {F}}}V(x)\leqslant 0 \end{aligned}$$
(A18)

for all \(x\in \mathbb {R}^{kn}\), which proves strong invariance of \({\mathcal {S}}_p(C)\) for all \(C>0\).

  1. 2)

    Let \(p=\infty\). Consider

    $$\begin{aligned} V(x) = \max _{i\in {\mathcal {I}}} \Vert x_i\Vert _\infty \end{aligned}$$
    (A19)

    as a Lyapunov function candidate. Since the proof shares the same structure and reasoning as the case \(p\in [1,\infty )\), we only provide a sketch of the proof.

In this case, the set \(\alpha (x)\) in (46) is

$$\begin{aligned} \alpha (x) = \{ i\in {\mathcal {I}}\,|\,\Vert x_i\Vert _\infty = V(x) \}, \end{aligned}$$
(A20)

whereas the generalized gradient of V reads

$$\begin{aligned} \partial V(x)=\mathrm {co} \{&e_i\otimes ({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )\,|\,e_i\in \mathbb {R}^n, \nonumber \\&e_\ell \in \mathbb {R}^k, |x_{i,\ell }|=V(x) \}. \end{aligned}$$
(A21)

For the case \(x\in {\mathcal {C}}\), we have \(\max {\mathcal {L}}_{{\mathcal {F}}[h]}V(x(t)) = 0 \leqslant 0\) by using the same argument as the case \(p\in [1,\infty )\). Hence, we omit the details.

For the case \(x\notin {\mathcal {C}}\), we first show that there exists an index \(i\in \alpha (x)\) such that \({\bar{L}}_ix \ne {\mathbf {0}}\) by using contradiction. If \(\alpha (x)={\mathcal {I}}\), the conclusion follows as the case (1). If \(\alpha (x)\subsetneq {\mathcal {I}}\) and \({\bar{L}}_ix = {\mathbf {0}}\) for all \(i\in \alpha (x)\), then for any \(i\in \alpha (x)\) with \(j\in N_i{\setminus }\alpha (x)\) we have

$$\begin{aligned} 0= ({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )^{\mathrm{T}}{\bar{L}}_ix \end{aligned}$$
(A22)
$$\begin{aligned}&\quad= \textstyle \sum \limits _{j\in N_i} ( |x_{i,\ell }| - {\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })x_{j,\ell } ) \\& \quad\geqslant \textstyle \sum \limits _{j\in N_i} (\Vert x_i\Vert _\infty - \Vert x_{j}\Vert _\infty ) \\&\quad> 0, \end{aligned}$$
(A23)

where \(e_i\otimes ({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )\in \partial V(x)\) and the last inequality is implied by \(\Vert x_i\Vert >\Vert x_j\Vert\) for any \(j\in N_i{\setminus }\alpha (x)\). This is a contradiction.

For the index \(i\in \alpha (x)\) satisfying \({\bar{L}}_ix\ne {\mathbf {0}}\), it follows from Assumption 1 that there exists \(\gamma >0\) such that for any \(\nu \in \bar{{\mathcal {F}}}(x)\), it holds that \(\nu _i = -\gamma {\bar{L}}_ix\). Using the same reasoning as in case (1), by choosing \(\zeta \in \partial V(x)\) as \(\zeta = e_i\otimes ({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )\) (recall that \(i\in \alpha (x)\)), it follows from (6) that

$$\begin{aligned} {\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x) = \{-\gamma ({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )^{\mathrm{T}}{\bar{L}}_ix\}. \end{aligned}$$
(A24)

Next, by observing (67), we have \(({\mathcal {F}}[{{\,\mathrm{sgn}\,}}](x_{i,\ell })e_\ell )^{\mathrm{T}}{\bar{L}}_ix \geqslant 0,\) which implies \({\mathcal {L}}_{\bar{{\mathcal {F}}}[h]}V(x)\subset \mathbb {R}_{\leqslant 0}\).

In summary, we have

$$\begin{aligned} \max {\mathcal {L}}_{{\mathcal {F}}}V(x)\leqslant 0 \end{aligned}$$
(A25)

for all \(x\in \mathbb {R}^{kn}\), which proves strong invariance of \({\mathcal {S}}_p(C)\) for all \(C>0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wei, J., Besselink, B., Wu, J. et al. Finite-time consensus protocols for multi-dimensional multi-agent systems. Control Theory Technol. 18, 419–430 (2020). https://doi.org/10.1007/s11768-020-00022-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11768-020-00022-y

Keywords

Navigation