Skip to main content
Log in

On time-triggered and event-based control of integrator systems over a shared communication system

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

When analyzing networked control systems, where the control loop is closed over a communication system, it is crucial to take the communication system into account. Hence, time-triggered and event-based control of an integrator system with noise over a shared communication system is analyzed. Thereby, analytical models of different communication systems are used and the analysis is focused on the effect of the communication system on the performance, as well as the interaction between control and communication. For time-triggered control, where the event times are known in advance, a deterministic communication protocol can be used. Hence, time-triggered control with the two most well-known deterministic communication protocols, time division multiple access (TDMA) and frequency division multiple access (FDMA), is analyzed. For event-based control, where the events appear at random times, a contention-based communication protocol should be used. Hence, event-based control is analyzed with different contention-based communication protocols: pure ALOHA, slotted ALOHA, a queueing system, and Erlang’s loss model. It turns out that time-triggered control with either TDMA or FDMA outperforms event-based control with pure or slotted ALOHA. However, event-based control with a properly designed queueing system gives an even better performance. Thus, we conclude that it is crucial to take the details of the communication system into account.

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. The vulnerable period is the time during which the sending of another agent leads to a collision.

  2. Self-interference means that a packet of an agent collides with another packet of the same agent.

References

  1. Åström KJ, Bernhardsson BM (2002) Comparison of Riemann and Lebesgue sampling for first order stochastic systems. In: Proceedings of the conference on decision and control (CDC), Las Vegas, Nevada, USA, pp 2011–2016

  2. Abramson N (1970) THE ALOHA SYSTEM—another alternative for computer communications. In: Proceedings of the fall joint computer conference, pp 281–286

  3. Abramson N (2009) The ALOHANet—surfing for wireless data. IEEE Commun Mag 47(12):21–25

    Article  Google Scholar 

  4. Aldous D (1987) Ultimate instability of exponential back-off protocol for acknowledgment-based transmission control of random access communication channels. IEEE Trans Inf Theory 33(2):219–223

    Article  MathSciNet  MATH  Google Scholar 

  5. Anta A, Tabuada P (2010) To sample or not to sample: self-triggered control for nonlinear systems. IEEE Trans Autom Control 55(9):2030–2042

    Article  MathSciNet  Google Scholar 

  6. Blind R, Allgöwer F (2011) Analysis of networked event-based control with a shared communication medium: part I—pure ALOHA. In: Proceedings of the 18th IFAC World Congress, Milano, Italy, pp 10092–10097

  7. Blind R, Allgöwer F (2011) Analysis of networked event-based control with a shared communication medium: part II—slotted ALOHA. In: Proceedings of the 18th IFAC World Congress, Milano, Italy, pp 8830–8835

  8. Blind R, Allgöwer F (2011) On the optimal sending rate for networked control systems with a shared communication medium. In: Proceedings of the conference on decision and control (CDC), Orlando, FL, USA, pp 4704–4709

  9. Brun O, Garcia JM (2000) Analytical solution of finite capacity M/D/1 queues. J Appl Probab 37(4):1092–1098

    Article  MathSciNet  MATH  Google Scholar 

  10. Cervin A, Henningsson T (2008) Scheduling of event-triggered controllers on a shared network. In: Proceedings of the conference on decision and control (CDC), Cancun, Mexico, pp 3601–3606

  11. Davis MHA, Howl JM (1997) A Markovian analysis of the finite-buffer M/D/1 queue. Proc R Soc Lond Ser A Math Phys Eng Sci 453(1964):1947–1962

    Article  MathSciNet  MATH  Google Scholar 

  12. Erlang AW (1917) Solution of some problems in the theory of probabilities of significance in automatic telephone exchanges. Post Off Electr Eng J 10:189–197

    Google Scholar 

  13. Feller W (1950) An introduction to probability theory and its applications, 1st edn. Wiley, New York

    MATH  Google Scholar 

  14. Heemels WPMH, Sandee JH (2008) Analysis of event-driven controllers for linear systems. Int J Control 81(4):571–590

    Article  MathSciNet  MATH  Google Scholar 

  15. Henningsson T, Cervin A (2010) A simple model for the interference between event-based control loops using a shared medium. In: Proceedings of the conference on decision and control (CDC), Atlanta, GA, USA, pp 3240–3245

  16. Henningsson T, Johannesson E, Cervin A (2008) Sporadic event-based control of first-order linear stochastic systems. Automatica 44(11):2890–2895

    Article  MathSciNet  MATH  Google Scholar 

  17. Heyman DP, Sobel MJ (1982) Stochastic models in operations research, vol I. McGraw-Hill Book Company, New York

    MATH  Google Scholar 

  18. Hristu-Varsakelis D, Kumar P (2002) Interrupt-based feedback control over a shared communication medium. In: Proceedings of the conference on decision and control (CDC), Las Vegas, Nevada, USA, pp 3223–3228

  19. Jacobson V (1988) Congestion avoidance and control. ACM SIGCOMM Comput Commun Rev 18(4):314–329

    Article  Google Scholar 

  20. Kelly F (1985) Stochastic models of computer communication systems. J R Stat Soc Ser B (Methodological) 47(3):379–395

    Google Scholar 

  21. Kleinrock L (1975) Queueing systems: volume I—theory. Wiley Insterscience, New York

    Google Scholar 

  22. Kleinrock L, Lam SS (1973) Packet-switching in a slotted satellite channel. National Computer Conference, AFIPS Conference Proceedings, New York, NY, USA, pp 703–710

  23. Kleinrock L, Tobagi F (1975) Packet switching in radio channels: part I—carrier sense multiple-access modes and their throughput-delay characteristics. IEEE Trans Commun 23(12):1400–1416

    Article  MATH  Google Scholar 

  24. Kofman E, Braslavsky JH (2006) Level crossing sampling in feedback stabilization under data-rate constraints. In: Proceedings of the conference on decision and control (CDC), San Diego, CA, USA, pp 4423–4428

  25. Lam S, Kleinrock L (1975) Packet switching in a multiaccess broadcast channel: dynamic control procedures. IEEE Trans Commun 23(9):891–904

    Article  MATH  Google Scholar 

  26. Lam SS, Kleinrock L (1975) Dynamic control schemes for a packet switched multi-access broadcast channel. National Computer Conference, AFIP Conference Proceedings, Anaheim, CA, USA, pp 143–153

  27. Liu X, Goldsmith A (2004) Wireless medium access control in networked control systems. In: Proceedings of the American Control Conference (ACC), Boston, MA, USA, pp 3605–3610

  28. Lunze J, Lehmann D (2010) A state-feedback approach to event-based control. Automatica 46(1):211–215

    Article  MathSciNet  Google Scholar 

  29. Mazo M, Tabuada P (2011) Decentralized event-triggered control over wireless sensor/actuator networks. IEEE Trans Autom Control 56(10):2456–2461

    Article  MathSciNet  Google Scholar 

  30. Mesquita AR, Hespanha JP, Nair GN (2012) Redundant data transmission in control/estimation over lossy networks. Automatica 48(8):1612–1620

    Article  MathSciNet  MATH  Google Scholar 

  31. Molin A, Hirche S (2011) Optimal design of decentralized event-triggered controllers for large-scale systems with contention-based communication. In: Proceedings of the conference on decision and control (CDC), Orlando, FL, USA, pp 4710–4716

  32. Navarro DJ, Fuss IG (2009) Fast and accurate calculations for first-passage times in Wiener diffusion models. J Math Psychol 53(4):222–230

    Article  MathSciNet  MATH  Google Scholar 

  33. Rabi M, Johansson KH (2009) Scheduling packets for event-triggered control. In: Proceedings of European Control Conference (ECC), Budapest, Hungary, pp 3779–3784

  34. Rabi M, Moustakides GV, Baras JS (2012) Adaptive sampling for linear state estimation. SIAM J Control Optim 50(2):672–702

    Article  MathSciNet  MATH  Google Scholar 

  35. Rabi M, Stabellini L (2008) Analysis of networked estimation under contention-based medium access. In: Proceedings of the IFAC World Congress, Seoul, Korea, pp 10283–10288

  36. Ramesh C, Sandberg H, Johansson KH (2011) Steady state performance analysis of multiple state-based schedulers with CSMA. in: Proceedings of the conference on decision and control (CDC), Orlando, FL, USA, pp 4729–4734

  37. Roberts L (1975) ALOHA packet system with and without slots and capture. ACM SIGCOMM Comput Commun Rev 5(2):28–42

    Article  Google Scholar 

  38. Rom R, Sidi M (1990) Multiple access protocols. Springer, New York

    Book  MATH  Google Scholar 

  39. Sant D (1980) Throughput of unslotted ALOHA channels with arbitray packet interarrival time distributions. IEEE Trans Commun 28(8):1422–1425

    Article  Google Scholar 

  40. Sommer J, Blind R (2007) Optimized resource dimensioning in an embedded CAN-CAN gateway. In: International symposium on industrial embedded systems, pp 55–62

  41. Tabuada P (2007) Event-triggered real-time scheduling of stabilizing control tasks. IEEE Trans Autom Control 52(9):1680–1685

    Article  MathSciNet  Google Scholar 

  42. Tanenbaum AS (2003) Computer networks. Pearson Education, New Jersey

    Google Scholar 

  43. Walsh GC, Ye H, Bushnell L (1999) Stability analysis of networked control systems. In: Proceedings of the American Control Conference (ACC), San Diego, CA, USA, pp 2876–2880

  44. Wang X, Lemmon MD (2008) Event-triggered broadcasting across distributed networked control systems. In: Proceedings of the American Control Conference (ACC), Seattle, WA, USA, pp 3139–3144

  45. Wang X, Lemmon MD (2011) Event-triggering in distributed networked control systems. IEEE Trans Autom Control 56(3):586–601

    Article  MathSciNet  Google Scholar 

  46. Feller William (1954) Diffusion processes in one dimension. Trans Am Math Soc 77(1):1–31

    Article  MathSciNet  MATH  Google Scholar 

  47. Zhang W (2003) Stabilization of networked control systems over a sharing link using ALOHA. In: Proceeding of the Conference on Decision and Control (CDC) (December), Maui, Hawaii USA, pp 204–209

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rainer Blind.

Additional information

This work was supported by the Deutsche Forschungsgemeinschaft (DFG) Priority Programme: Control Theory of Digitally Networked Dynamical Systems. Parts of this work were presented at the 18th IFAC World Congress, Milano, Italy, 2011 and the 50th Conference on Decision and Control (CDC), Orlando, FL, USA, 2011.

Appendices

Appendix A: Remaining proofs of Sect. 3

In [1], the expected cost and interevent time for controlling system (2) with an impulsive input is studied for \(\sigma _{i} = 1\) and an ideal communication. In [33], this result is extended to the case of a communication system with random packet loss. To prove Theorems 4, 7, and 8, we have to extend these results to arbitrary \(\sigma _{i}\) and packet delay.

For simplicity of notation, we drop the subindex which indicates the agent.

1.1 Expected time between two successful transmissions

Lemma 35

Suppose, the expected time between two transmission attempts is \(T\) and the packet loss probability is \(p\). Then the expected time between two successful transmissions \(T_{S}\) is

$$\begin{aligned} {{\mathrm{E}}}[ T_{S} ] = \frac{T}{1-p}. \end{aligned}$$

Proof

$$\begin{aligned} {{\mathrm{E}}}[ T_{S} ] = \sum _{k=1}^{\infty } (1-p) p^{k-1} k T = T \frac{(1-p)}{p} \sum _{k=0}^{\infty } k p^{k} = \frac{T}{1-p} \end{aligned}$$

\(\square \)

1.2 Additional cost due to delay

Time-triggered and event-based control with packet loss has been studied in [33]. In the considered setup packets are also delayed, so this effect must be taken into account. To show how delayed events affect the cost, we use \(\tilde{x}\) to denote the state of the system with delayed events, whereas \(x\) remains the state of the system without delayed events. Figure 10 shows an example of this case. Suppose an event is generated at time \(t_{k}\). If the corresponding impulse is applied immediately, then the state will be reset to the origin. Due to the delay \(d_{k}\), the impulse is applied at time \(t_{k} + d_{k}\). Since the process continues in the meantime, the impulse will not reset the state to the origin. Thus, the cost of the system with delayed events will be different from the cost of the system without delayed events as stated in the following lemma.

Fig. 10
figure 10

The effect of delayed events

Lemma 36

Suppose input (3) is used to control system (2). Then, the cost of the closed loop system with delayed events is \(J + \sigma ^{2} d\), where \(J\) is the cost of the system without delay and \(d\) the expected delay.

Proof

Since the interevent times are iid, it suffices to integrate between the two successful packet transmissions \(t_k\) and \(t_{k+1}\). Thus, we get for the cost of the process with delayed events:

$$\begin{aligned} \tilde{J} = \frac{{{\mathrm{E}}}\left[ \int _{t_k}^{t_{k+1}} \tilde{x}(t)^2\,\mathrm{d}t\right] }{{{\mathrm{E}}}[t_{k+1}-t_{k}]}. \end{aligned}$$
(68)

Obviously, we get for the denominator \({{\mathrm{E}}}[t_{k+1} - t_{k}] = {{\mathrm{E}}}[T_{S}]\). Thus, we continue as follows:

$$\begin{aligned} \tilde{J}&= \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_{k+1}} \tilde{x}(t)^2\,\mathrm{d}t\right] \\&= \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_k+d_{k}} \tilde{x}(t)^2\,\mathrm{d}t\right] + \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k+d_{k}}^{t_{k+1}} \tilde{x}(t)^2\,\mathrm{d}t\right] \\&= \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_k+d_{k}} \left( x(t)+x(t_k)\right) ^2 \,\mathrm{d}t\right] + \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k+d_{k}}^{t_{k+1}} x(t)^2\,\mathrm{d}t\right] \\&= \underbrace{ \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_{k+1}} x(t)^2\,\mathrm{d}t\right] }_{J} + \underbrace{ \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_k+d_{k}} x(t_k)^2\,\mathrm{d}t\right] }_{J_{d}}\\&\quad + \underbrace{\frac{2}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}\left[ \int _{t_k}^{t_k+d_{k}} x(t) x(t_k)\,\mathrm{d}t\right] }_{= J^*}. \end{aligned}$$

Note that \(J^* = 0\) because \(x(t)\) and \(x(t_{k})\) are independent and \({{\mathrm{E}}}[x(t)] = 0\). Thus, it remains to check \(J_{d} = \frac{1}{{{\mathrm{E}}}[T_{S}]} \int _{t_k}^{t_k+d_{k}} {{\mathrm{E}}}[x(t_k)^2]\,\mathrm{d}t\). Since the delay \(d_{k}\) is independent from \(x(t_{k})\) and \({{\mathrm{E}}}[d_{k}] = d\), we get for the additional cost due to delay \(J_{d} = \frac{1}{{{\mathrm{E}}}[T_{S}]} {{\mathrm{E}}}[x(t_k)^2]d\). Since the process was reset to the origin at the previous successful transmission \(t_{k-1}\), we have \({{\mathrm{E}}}[x(t_{k})^{2}] = \sigma ^{2} {{\mathrm{E}}}[T_{S}]\) and consequently \(J_{d} = \sigma ^{2 }d\). \(\square \)

1.3 Proof of Theorem 4

To prove Theorem 4, we first extend the cost of time-triggered control as given in [1] to an arbitrary noise intensity \(\sigma \). Based on this result, we then derive the cost of time-triggered control with packet loss.

Lemma 37

Suppose, system (2) is controlled by an impulsive time-triggered control scheme with sampling time \(T_{\mathrm{TT }}\) and an ideal communication without loss and delay. Then the cost is

$$\begin{aligned} J_{\mathrm{TT }} = \sigma ^{2} \frac{T_{\mathrm{TT }}}{2}. \end{aligned}$$
(69)

Proof

$$\begin{aligned} J_{\mathrm{TT }} = \frac{1}{T_{\mathrm{TT }}} {{\mathrm{E}}}\left[ \int _{0}^{T_{\mathrm{TT }}} x^{2}(t)\,{\hbox {d}}t\right] = \frac{1}{T_{\mathrm{TT }}} \int _{0}^{T_{\mathrm{TT }}} {{\mathrm{E}}}[x^{2}(t)]\,{\hbox {d}}t = \frac{1}{T_{\mathrm{TT }}} \int _{0}^{T_{\mathrm{TT }}} \sigma ^{2}t\,{\hbox {d}}t=\sigma ^{2} \frac{T_{\mathrm{TT }}}{2}. \end{aligned}$$

\(\square \)

Lemma 38

Suppose, system (2) is controlled by an impulsive time-triggered control scheme with sampling time \(T_{\mathrm{TT }}\) and a packet loss probability \(p\). Then the cost is

$$\begin{aligned} J_{\mathrm{TT }} = \sigma ^{2} \left( \frac{T_{\mathrm{TT }}}{2} + \frac{T_{\mathrm{TT }} p}{(1 - p)} \right) . \end{aligned}$$

Proof

To prove Lemma 38, we use \(J_{m}\) to denote the cost for the case that the sampling time is extended to \(m T_{\mathrm{TT }}\). From Lemma 37, we have

$$\begin{aligned} J_{m} = \frac{{{\mathrm{E}}}\left[ \int _{0}^{m T_{\mathrm{TT }}} x^{2}(t)\,{\hbox {d}}t \right] }{m T_{\mathrm{TT }}} = \sigma ^{2} \frac{m T_{\mathrm{TT }}}{2} \end{aligned}$$

and thus,

$$\begin{aligned} {{\mathrm{E}}}\left[ \int _{0}^{m T_{\mathrm{TT }}} x^{2}(t)\,{\hbox {d}}t \right] = J_{m} m T_{\mathrm{TT }} = \sigma ^{2} \frac{m^{2} T_{\mathrm{TT }}^{2}}{2}. \end{aligned}$$

Now, we consider the term \({{\mathrm{E}}}[ \int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t]\), where \(T_{S}\) is the expected time between two successful packet transmissions.

$$\begin{aligned} {{\mathrm{E}}}\left[ \int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t\right]&= \sum _{m=1}^{\infty } (1-p) p^{m-1} {{\mathrm{E}}}\left[ \int _{0}^{m T_{\mathrm{TT }}} x^{2}(t)\,{\hbox {d}}t \right] \\&= \frac{\sigma ^{2} T_{\mathrm{TT }}^{2}}{2} \frac{1-p}{p} \sum _{m=0}^{\infty } m^{2} p^{m}= \frac{\sigma ^{2} T_{\mathrm{TT }}^{2}}{2} \frac{1+p}{(1-p)^{2}}. \end{aligned}$$

Thus, we get for the cost

$$\begin{aligned} J_{\mathrm{TT }} = \frac{{{\mathrm{E}}}\left[ \int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t \right] }{{{\mathrm{E}}}[T_{S}]} = \frac{\sigma ^{2} T_{\mathrm{TT }}}{2} \frac{1+p}{1-p} = \sigma ^{2} \left( \frac{T_{\mathrm{TT }}}{2} + \frac{T_{\mathrm{TT }} p}{(1 - p)} \right) . \end{aligned}$$

\(\square \)

Proof of Theorem 4

Theorem 4 follows from combining Lemmas 36 and 38. \(\square \)

1.4 Proof of Theorem 7

To prove Theorem 7, we use a result from [46].

Lemma 39

([46]) Suppose, the lower threshold is \(\underline{\Delta }\) and the upper threshold is \(\overline{\Delta }\) and system (2) is started at \(x_{0}\); \(\underline{\Delta } < x_{0} < \overline{\Delta }\). Then, the expected time to reach one of these thresholds is characterized as the unique solution of

$$\begin{aligned} \frac{1}{2} \sigma ^{2} \frac{\partial ^{2}}{\partial x^{2}} \Psi (x_{0}) = -1 \qquad \text {with }\Psi (\underline{\Delta }) = \Psi (\overline{\Delta }) = 0. \end{aligned}$$
(70)

Proof of Theorem 7

Proof of Theorem 7 For the considered event-based control, we have \(\underline{\Delta } = - \Delta \), \(\overline{\Delta } = + \Delta \). Thus, the solution of (70) is

$$\begin{aligned} \Psi (x_{0}) = \frac{1}{\sigma ^{2}} (\Delta ^{2} - x_{0}). \end{aligned}$$

Now, note that in the considered setup, the system is reset to the origin, i.e., \(x_{0} = 0\). Thus, the expected interevent time is

$$\begin{aligned} T_{\mathrm{EB }} = \Psi (0) = \frac{\Delta ^{2}}{\sigma ^{2}}. \end{aligned}$$

\(\square \)

1.5 Proof of Theorem 8

Again, we extend the cost given in [1] to the case of an arbitrary noise intensity \(\sigma \). Based on this result, we derive the cost of event-based control with packet loss similar to [33].

Lemma 40

Suppose, system (2) is controlled by an impulsive event-based control scheme with threshold increment \(\Delta \) and an ideal communication without loss and delay. Then the cost is

$$\begin{aligned} J_{\mathrm{EB }} = \sigma ^{2} \frac{T_{\mathrm{EB }}}{6}. \end{aligned}$$
(71)

Proof

The distribution of the state follows from the steady state of the Kolmogorov forward equation

$$\begin{aligned} 0 = \frac{1}{2} \frac{\partial ^{2}}{\partial x^{2}} \sigma ^{2} f(x) \qquad \text {with }\int _{-\Delta }^{\Delta } f(x)\,{\hbox {d}}x = 1, f(-\Delta ) = f(\Delta ) = 0. \end{aligned}$$

This equation has the solution

$$\begin{aligned} f(x) = \frac{\Delta - |x|}{\Delta ^{2}}. \end{aligned}$$

Consequently, the variance is

$$\begin{aligned} J_{\mathrm{EB }} = \int _{-\Delta }^{\Delta } x^{2} f(x)\,{\hbox {d}}x= \frac{\Delta ^{2}}{6} = \sigma ^{2} \frac{T_{\mathrm{EB }}}{6}. \end{aligned}$$

\(\square \)

Lemma 41

Suppose, system (2) is controlled by an impulsive event-based control scheme with threshold increment \(\Delta \) and a packet loss probability \(p\). Then the cost is

$$\begin{aligned} J_{\mathrm{EB }} = \sigma ^{2} \left( \frac{T_{\mathrm{EB }}}{6} + \frac{T_{\mathrm{EB }} p}{(1 - p)}\right) . \end{aligned}$$
(72)

Proof

To prove Lemma 41, we use \(T_{m}\) to denote the time between \(m\) transmission attempts and \(T_{S}\) the time between two successful transmissions. We start by considering the term \({{\mathrm{E}}}[\int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t ]\)

$$\begin{aligned} {{\mathrm{E}}}\left[ \int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t \right]&= \sum _{m=1}^{\infty } (1-p) p^{m-1} {{\mathrm{E}}}\left[ \int _{0}^{T_{m}} x^{2}(t)\,{\hbox {d}}t\right] \\&=\frac{1-p}{p} \sum _{m=0}^{\infty } p^{m} \sum _{n=1}^{m} {{\mathrm{E}}}\left[ \int _{T_{n-1}}^{T_{n}} x^{2}(t)\,{\hbox {d}}t\right] . \end{aligned}$$

To proceed, we consider the term \(\nu _{n} := {{\mathrm{E}}}[\int _{T_{n-1}}^{T_{n}} x^{2}(t)\,{\hbox {d}}t]\).

$$\begin{aligned} \nu _{n} := {{\mathrm{E}}}\left[ \int _{T_{n-1}}^{T_{n}} x^{2}(t)\,{\hbox {d}}t\right] = {{\mathrm{E}}}\left[ x^{2}(T_{n-1}) \int _{T_{n-1}}^{T_{n}}\,{\hbox {d}}t + \int _{T_{n-1}}^{T_{n}} (x(t) - x(T_{n-1}))^{2}\,{\hbox {d}}t\right] . \end{aligned}$$

Now, note that \(x(T_{n})\) is always an integer multiple of \(\Delta \) and similar to the random process \(\sum _{m=1}^{n} \theta (m) \Delta \), where \(\theta (m) \in \{-1, 1\}\) with and . Thus, \(\nu _{n}\) becomes

$$\begin{aligned} \nu _{n} = {{\mathrm{E}}}\left[ \left( \sum _{m=1}^{n} \theta _{m} \Delta \right) ^{2}\right] {{\mathrm{E}}}[T_{\mathrm{EB }}] + {{\mathrm{E}}}\left[ \int _{0}^{T_{\mathrm{EB }}} x^{2}(t)\,{\hbox {d}}t |x(0) = 0\right] = (n-1) \frac{\Delta ^{4}}{\sigma ^{2}} + \frac{\Delta ^{4}}{6 \sigma ^{2}}. \end{aligned}$$

Now, we proceed with the term \({{\mathrm{E}}}[\int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t]\)

$$\begin{aligned} {{\mathrm{E}}}\left[ \int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t\right]&= \frac{1-p}{p} \sum _{m=0}^{\infty } p^{m} \sum _{n=1}^{m} \nu _{n} \\&= \frac{1-p}{p} \sum _{m=0}^{\infty } p^{m} \sum _{n=1}^{m} (n-1) \frac{\Delta ^{4}}{\sigma ^{2}} + \frac{\Delta ^{4}}{6 \sigma ^{2}} \\&= \frac{1-p}{p} \frac{\Delta ^{4}}{\sigma ^{2}} \sum _{m=0}^{\infty } p^{m} \left( \frac{m(m-1)}{2} + \frac{m}{6} \right) \\&= \frac{\Delta ^{4}}{\sigma ^{2}} \left( \frac{p}{(1-p)^{2}} + \frac{1}{6 (1-p)} \right) . \end{aligned}$$

Finally, Lemma 41 follows from \(J_{\mathrm{EB }, i} = \frac{{{\mathrm{E}}}[\int _{0}^{T_{S}} x^{2}(t)\,{\hbox {d}}t ]}{{{\mathrm{E}}}[T_{S}]}\). \(\square \)

Proof of Theorem 8

Theorem 8 follows from combining Lemmas 36 and 41. \(\square \)

1.6 Proof of Theorem 12

To show that the arrival process of event-based control converges to a Poisson process for \(N \rightarrow \infty \), we use the Palm–Khintchine Theorem and the following two assumptions, definition and short discussion, from [17].

Assumption 42

For all \(N\) sufficiently large,

$$\begin{aligned} \lambda _{1, N} + \cdots + \lambda _{N, N} = \lambda _{\Sigma } < \infty , \end{aligned}$$
(73)

where \(\lambda _{j, N}\) is the sending rate of agent \(j\) for the case that there are \(N\) agents.

Assumption 43

Given \(\epsilon > 0\), for each \(t > 0\) and \(N\) sufficiently large,

$$\begin{aligned} F_{j, N}(t) \le \epsilon , \qquad j = 1, \cdots , N, \end{aligned}$$
(74)

where \(F_{j, N}\) is the CDF of agent \(j\) for the case that there are \(N\) agents.

Definition 44

For each \(N\) define

$$\begin{aligned} L_{0, N}(t) = L_{1, N}(t) + \cdots + L_{N, N}(t), \end{aligned}$$
(75)

where \(L_{j, N}(t)\) is a stochastic process, which counts the number of events of agent \(j\) that occur by time \(t\) for the case that there are \(N\) agents.

Assumption 43 asserts that as \(N\) increases, the processes being combined have renewals very infrequently. Assumption 42 shows that \(L_{0, N+1}(t)\) is not formed by adding another process to \(L_{0, N}(t)\). As \(N\) increases, the processes being combined are changed so that (at least for large \(N\)) the asymptotic rate at which renewals occur is a constant.

Theorem 45

(Palm–Khintchine) [17, Theorem 5.15] Under Assumptions 42 and 43, as \(N \rightarrow \infty \), \(\{L_{0, N}(t); t \ge 0\}\) approaches a Poisson process.

Proof of Theorem 12

Since Assumption 42 is part of Theorem 12, it remains to show that Assumption 43 holds. From the scaling property and the definition of the CDF, it follows that . Since \(f(x |1)\) is continuous, if follows from the mean value theorem that for some \(z \in (0,t)\). Moreover, since for \(N \rightarrow \infty \), we see that for each \(\epsilon > 0 \) and \(t > 0\) there exists an \(N\) such that \(F_{j, N}(t) \le \epsilon \), i.e., Assumption 43 holds. \(\square \)

Appendix B: Remaining proofs of Sect. 4

1.1 Proof of Theorem 13

For simplicity of notation, we drop the subindex indicating the agent and use it to indicate the scheme instead. Moreover, we use \(\tilde{p}\) to denote the probability of an unsuccessful transmission and \(\tilde{d}\) for the delay between a packet transmission and its reception.

When using Scheme 1, all packets that are not successfully transmitted are lost and the delay is only the delay between event generation and the reception of the corresponding packet. Thus, for Scheme 1, the additional cost due to loss and delay is

$$\begin{aligned} \frac{\tilde{p}_{1}}{1-\tilde{p}_{1}} T + \tilde{d}_{1}. \end{aligned}$$
(76)

When using Scheme 2, all packets that are not transmitted successfully are retransmitted at some later time. Thus, there is no packet loss but an additional delay due to the retransmissions.

$$\begin{aligned} d_{2}&= \tilde{d}_{2} + 0 T (1-\tilde{p}_{2}) + 1 T (1-\tilde{p}_{2}) \tilde{p}_{2} + 2 T (1-\tilde{p}_{2}) \tilde{p}_{2}^{2} + 3 T (1-\tilde{p}_{2}) \tilde{p}_{2}^{3} \\&\quad + \cdots =\tilde{d}_{2} + (1-\tilde{p}_{2}) \tilde{p}_{2} T \sum _{m=0}^{\infty } m \tilde{p}_{2}^{m-1} = \tilde{d}_{2} + \frac{\tilde{p}_{2}}{1-\tilde{p}_{2}} T. \end{aligned}$$

Thus, for Scheme 2, the additional cost due to delay becomes

$$\begin{aligned} \frac{\tilde{p}_{2}}{1 - \tilde{p}_{2}} T + \tilde{d}_{2}. \end{aligned}$$
(77)

Now, we can compare the additional cost due to loss and delay of Scheme 1 and Scheme 2, given by (76) and (77). Due to the retransmission of packets, the load when using Scheme 2 is larger than the load when using Scheme 1. Since we assumed that either the probability of an unsuccessful transmission or delay, or both, is strictly increasing with the network load, Scheme 1 gives a better performance than Scheme 2. \(\square \)

1.2 Proof of Lemma 17

We start by bounding the probability of self-interference, i.e., the term \((1-\int _{0}^{\rho _{i}} f_{i}(x |1)\,{\hbox {d}}t )^{2}\). Since \(f_{i}(t |1)\) is a PDF, \(0 \le 1-\int _{0}^{x} f_{i}(t |1)\,{\hbox {d}}t \le 1\) holds. From the assumption \(f_{i}(t |1) \le 1\), we also have \(1 - x \le 1 - \int _{0}^{x} f_{i}(t |1)\,{\hbox {d}}t\), i.e.,

$$\begin{aligned} \max \{0, 1-x\} \le 1 - \int _{0}^{x} f_{i}(t |1)\,{\hbox {d}}t \le 1. \end{aligned}$$
(78)

Since \((1-\rho _{i})^{2} = 1 - 2 \rho _{i} + \rho _{i}^{2} \le 1 - 2 \rho _{i}\), we can bound the probability of self-interference as follows

$$\begin{aligned} b_{p}(\rho _{i}) \le \left( 1-\int _{0}^{\rho _{i}} f_{i}(x |1)\,{\hbox {d}}t \right) ^{2} \le 1, \end{aligned}$$
(79)

with \(b_{p}(\rho )\) as defined in (24).

Now, we continue with the probability of a collision with another user, i.e., the term \(p_{co}(\rho _{j}) := 1 - \int _{0}^{2\rho _{j}} [1-\int _{0}^{x}f_j(t |1)\,{\hbox {d}}t]\,{\hbox {d}}x\). Since \(p_{co}(\rho _{j})\) is a probability, we immediately have \(0 \le p_{co}(\rho _{j}) \le 1\). Moreover, from (78) we get \(1 - 2 \rho _{j} \le p_{co}(\rho _{j}) \le 1 - 2 \rho _{j} + 2 \rho _{j}^{2}\). Since \(p_{co}(\rho _{j})\) is increasing with \(\rho _{j}\) and the upper bound \(1 - 2 \rho _{j} + 2 \rho _{j}^{2}\) has a minimum at , we use \(a_{p}(\rho _{j})\) as upper bound for \(p_{co}(\rho _{j})\) for . Therefore, we get

$$\begin{aligned} b_{p}(\rho _{j}) \le 1 - \int _{0}^{2 \rho _{j}}\left[ 1 - \int _{0}^{x} f_{j}(t |1)\,{\hbox {d}}t\right] \,{\hbox {d}}x \le a_{p}(\rho _{j}), \end{aligned}$$
(80)

with \(a_{p}(\rho )\) and \(b_{p}(\rho )\) as defined in (23) and (24), respectively.

Finally, (22) follows using (79) and (80) in (21). \(\square \)

1.3 Proof of Theorem 20

We start by proving the case without self-interference. Equation (29) follows from (20) due to the fact that the vulnerable period is \(\tau \) for slotted ALOHA, instead of \(2 \tau \) for pure ALOHA.

Next, we analyze the losses due to self-interference. Figure 11 shows an event generation within a slot and the corresponding times. The time interval between the event generation and the end of the slot is called  forward recurrence time \(\tau _F\). Similarly, the time interval between the event generation and the begin of the slot is denoted backward recurrence time \(\tau _B\). As already stated, a self-interference occurs if the same user generates a new event while its packet is still waiting for the start of the next slot, i.e., during \(\tau _F\). Thus, the steady-state probability that a packet is not lost due to self-interference is

$$\begin{aligned} {{\mathrm{E}}}\left[ \left( 1 - \int _{0}^{\tau _F} f_i(t |\lambda _{i})\,{\hbox {d}}t \right) \right] = \frac{1}{\tau } \int _{0}^{\tau } \left( 1 - \int _{0}^{\tau -x} f_i(t |\lambda _{i})\,{\hbox {d}}t \right) \,{\hbox {d}}x. \end{aligned}$$
(81)

Finally, (28) follow from the fact that a packet of user \(i\) is lost if it interferes with a packet of user \(i\) or with a packet of any other user. \(\square \)

Fig. 11
figure 11

Forward and backward recurrence time

1.4 Proof of Lemma 23

First, we look at the losses due to self-interference. Since the corresponding term is a probability and the assumption \(f_{i}(x |1) \le 1\), we get

$$\begin{aligned} b_{s}(\rho ) \le \frac{1}{\rho _{i}} \int _{0}^{\rho _{i}} \left( 1 - \int _{0}^{\rho _{i}-x} f(t |1)\,{\hbox {d}}t \right) \,{\hbox {d}}x \le 1. \end{aligned}$$
(82)

By following the same arguments as in the proof of Lemma 17, we get

$$\begin{aligned} b_{s}(\rho _{j}) \le 1 - \int _{0}^{ \rho _{j}} \left[ 1 - \int _{0}^{x} f_{j}(t |1)\,{\hbox {d}}t\right] \,{\hbox {d}}x \le a_{s}(\rho _{j}). \end{aligned}$$
(83)

Finally, (31) follows using (82) and (83) in (30). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Blind, R., Allgöwer, F. On time-triggered and event-based control of integrator systems over a shared communication system. Math. Control Signals Syst. 25, 517–557 (2013). https://doi.org/10.1007/s00498-013-0117-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00498-013-0117-3

Keywords

Navigation