Skip to main content
Log in

Closed-Loop Control Strategy for Simulated Smoke Concentration in Aircraft Cargo Compartment Mock-Up

  • Manuscript
  • Published:
Fire Technology Aims and scope Submit manuscript

Abstract

In airworthiness verification flight test on fire smoke detection systems in aircraft cargo compartments, simulated smoke from a smoke generator that can replicate the characteristics of actual fire smoke should be used. In current studies, most of the boundary conditions of smoke generators are adjusted by open-loop control to fulfill the equivalence of the actual fire smoke. However, this method consumes a large amount of resources and has difficulty achieving optimal results from the complex turbulent flow field of smoke in an aircraft cargo compartment. To solve this problem, a computational fluid dynamics (CFD) model was first established to simulate smoke in the full-scale cargo compartment mock-up. Then, a state-space model with the exit parameters of the smoke generator as the input and the smoke concentration in the full-scale cargo compartment mock-up as the state variable was constructed using the data generated by the CFD model. Subsequently, a closed-loop simulated smoke concentration control strategy is proposed. In this strategy, the sensor in the full-scale cargo compartment mock-up collects the simulated smoke concentration signal and feeds it back to the controller. The controller processes the control target and feedback signal in real time. It realizes the automatic closed-loop control of the simulated smoke concentration by automatically adjusting the boundary condition parameters of the smoke generator. In the control law design, the control input constraint and the error between the linear state-space model and the real flow are considered, making the linear state-space model a nonlinear model. Finally, we prove the stability of the control system. The simulation shows that the designed control law can realize the closed-loop control of smoke concentration. The findings of this study can expand the control of simulated smoke flow field in airworthiness verification flight test from open loop to closed loop, and offer a new theoretical approach for smoke simulation technology.

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.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13

Similar content being viewed by others

References

  1. Blake D (2000) Aircraft cargo compartment smoke detector alarm incidents on US registered aircraft, 1974–1999. Federal Aviation Administration, Washington

    Google Scholar 

  2. Federal Aviation Administration (2014) Airworthiness standards: transport category airplanes part 25.858. GPO’s Federal Digital System, New York

    Google Scholar 

  3. Advisory Circular. AC 25-9A (1994) Smoke detection, penetration, evacuation tests and related flight manual emergency procedures. Federal Aviation Administration, Washington

    Google Scholar 

  4. Suo-Anttila J, Gill W, Gritzo LA, Blake D (2005) An evaluation of actual and simulated smoke properties. Fire Mater 29(2):91–107

    Article  Google Scholar 

  5. Suo-Anttila J, Gill W, Gritzo L (2003) Comparison of actual and simulated smoke for the certification of smoke detectors in aircraft cargo compartments, DOT/FAA/AR-03/34. Federal Aviation Administration, Washington

    Google Scholar 

  6. Behle K (2006) Determination of smoke quantities to be used for smoke detection performance ground and flight tests. In: 25th Congress of the International Council of the Aeronautical Sciences 2006, Hamburg, Germany. Curran Associates Inc

  7. Chen X, Shao Z, Yang J (2020) Comparison between actual and simulated smoke for smoke detection certification in aircraft cargo compartments using the CFD method. Fire Technol 56(2):469–488

    Article  Google Scholar 

  8. Zhao G, Wang LL (2014) Using helium smoke as a surrogate of fire smoke for the study of atrium smoke filling. J Fire Sci 32(5):431–447

    Article  Google Scholar 

  9. Wang LL, Zhao G (2013) Numerical study on smoke movement driven by pure helium in atria. Fire Saf J 61:45–53

    Article  Google Scholar 

  10. Karp M (2018) Smoke generator standardization and the effects of environmental variations. Federal Aviation Administration, Washington

    Google Scholar 

  11. Karp M (2019) Methods for characterizing theatrical smoke for smoke detection certification. Federal Aviation Administration, Washington

    Google Scholar 

  12. Freiling A (2018) Smoke detector response time comparison to artificial smoke sources. Federal Aviation Administration, Washington

    Google Scholar 

  13. Freiling A (2018) Smoke generator performance assessment. Federal Aviation Administration, Washington

    Google Scholar 

  14. Freiling A (2017) Standardization of smoke generation for certification testing. In: 2nd Airbus fire protection symposium, Atlantic City

  15. Emami T (2018) Characterization of smoke machines in testing aircraft smoke detectors. DOT/FAA/TC-TT18/9. Federal Aviation Administration, Washington

    Google Scholar 

  16. Blake D, Suo-Anttila J (2008) Aircraft cargo compartment fire detection and smoke transport modeling. Fire Saf J 43(8):576–582

    Article  Google Scholar 

  17. Oztekin ES, Blake D, Lyon RE (2013) Fire induced flow behavior in a ventilated aircraft cargo compartment. In: 13th international conference and exhibition on fire and materials 2013. Interscience Communications Ltd, San Francisco

  18. Oztekin ES (2014) Heat and mass transfer due to a small-fire in an aircraft cargo compartment. Int J Heat Mass Transf 73(1):562–573

    Article  MathSciNet  Google Scholar 

  19. Suo-Anttila J, Gill W, Gallegos C, Nelsen J (2003) Computational fluid dynamics code for smoke transport during an aircraft cargo compartment fire: transport solver, graphical user interface, and preliminary baseline validation, DOT/FAA/AR-03/49. Federal Aviation Administration, Washington

    Google Scholar 

  20. Suo-Anttila J, Gill W, Gallegos C, Nelsen J (2007) Cargo compartment smoke transport computational fluid dynamics code validation. Federal Aviation Administration, Washington

    Google Scholar 

  21. Freiling A (2020) Horizontal smoke velocity characterization for artificial smoke generators. Federal Aviation Administration, Washington

    Google Scholar 

  22. Karp M (2021) Smoke generator handbook for cargo smoke detection certification. Federal Aviation Administration, Washington

    Google Scholar 

  23. Chen XY, Shao ZY, Yang JZ (2020) Research on equivalence strategy of aircraft cargo compartment smoke based on inverse design. Acta Aeronaut Astronaut Sin 41(12):123991. https://doi.org/10.7527/S1000-68932020.23991

    Article  Google Scholar 

  24. Bhavithavya KK, MacKunis W, Drakunov SV, Golubev V (2020) Sliding mode estimation and closed-loop active flow control under actuator uncertainty. Int J Robust Nonlinear Control 30(16):6645–6640

    Article  MathSciNet  Google Scholar 

  25. Ren K, Chen Y, Gao CQ, Zhang WW (2020) Adaptive control of transonic buffet flows over an airfoil. Phys Fluids 32(9):96–106

    Article  Google Scholar 

  26. Taira K, Hemati M, Brunton SL et al (2020) Modal analysis of fluid flows: applications and outlook. AIAA J 58(3):998–1022

    Article  Google Scholar 

  27. Jayaprakash AK, Mackunis W (2022) Robust nonlinear control of fluid flow fields using a barrier function-based sliding mode control approach. In: 2022 IEEE conference on control technology and applications. pp 63-68

  28. McGrattan K, McDermott R, Hostikka S et al (2019) Fire dynamics simulator user’s guide. NIST special publication 1019, 6th edn. National Institute of Standards and Technology, Gaithersburg

    Google Scholar 

  29. Mulholland GW, Croarkin C (2000) Specific extinction coefficient of flame generated smoke. Fire Mater 24(1):227–230

    Article  Google Scholar 

  30. Krishnan SS, Lin KC, Faeth GM (2001) Extinction and scattering properties of soot emitted from buoyant turbulent diffusion flames. J Heat Transf 123(1):331

    Article  Google Scholar 

  31. Cunningham WJ (1962) An introduction to Lyapunov’s second method. Trans Am Inst Electr Eng II 80:325

    Google Scholar 

  32. Tu XH, Lai JF, Zhao K (2016) Robust adaptive control of nonlinear systems with asymmetric non-smooth saturation. In: 2016 Chinese control and decision conference (CCDC). Proceedings. pp 500–505

Download references

Acknowledgements

This study was supported by the multi-investment fund project of Tianjin Science and Technology Bureau of China under Project No. 21JCQNJC00940.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xiyuan Chen.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Control Law Design and Mathematical Derivation of the Stability Proof for the Closed-Loop Control of Simulated Smoke Concentration

Appendix: Control Law Design and Mathematical Derivation of the Stability Proof for the Closed-Loop Control of Simulated Smoke Concentration

1.1 Control Law Design

After the model identification is completed, we first transform the state-space model of simulated smoke obtained from the identification in Sect. 2.3 into a controllable standard type, which can be written as follows:

$$\begin{array}{*{20}l} \dot{x}_{1} = x_{2} \hfill \\ \dot{x}_{2} = x_{3} \hfill \\ \dot{x}_{3} = \varphi (x) + u \hfill \\ y = x_{1} , \hfill \\ \end{array}$$
(11)
$${\text{where}}{\kern 1pt} \,\varphi (x) = - b_{0} x_{1} - b_{1} x_{2} .$$
(12)

\(b_{0}\), \(b_{1}\) are all positive constants. The state variable \(x_{1}\) represents the smoke concentration in the full-scale cargo compartment and is also the system output. The variable \(u\) represents the system input, which is the outlet flow rate of the smoke generator.

Here, we consider the input of the control system to be limited and the flow rate of the smoke generator changes in a limited scope. Therefore, we used a hard saturation function \(H(u)\) for limited control input (Fig. 14). As shown in Equation (13), \(u_{m1}\) and \(u_{m2}\) are the two saturated inflection points of the control input signal \(u(t)\), which are positive constant. The parameters \(\overline{\delta } > 0\) and \(- \underline{\delta } < 0\) are the upper and lower bounds of the control input restricted hard saturation function, respectively. The slope \(k(t)\) is a bounded time-varying function, where we define the greater of \(\overline{\delta }\) and \(\underline{\delta }\) as \(\Psi = \max \left\{ {\overline{\delta } ,\underline{\delta } } \right\}\). As the outlet flow of the smoke generator cannot be negative, we set \(\underline{\delta }\), \(\overline{\delta }\) and \(k(t)\) as 0 g/s, 10 g/s, and 1 to simplify the calculation.

$$H(u) = \left\{ {\begin{array}{*{20}l} {\overline{\delta } ,u > u_{m1} } \hfill \\ {k(t)u, - u_{m2} \le u \le u_{m1} } \hfill \\ { - \underline{\delta } ,u < - u_{m2} } \hfill \\ \end{array} } \right..$$
(13)
Figure 14
figure 14

Asymmetric saturation function H (u) with dead zones and its approximation function S (u)

Because \(H(u)\) is an asymmetric and non-smooth hard saturation function, which increases the difficulty of system control law design, we choose a soft saturation function \(S(u)\) to approximate the asymmetric non-smooth hard saturation function \(H(u)\). This soft saturation function \(S(u)\) is expressed as follows:

$$S(u) = \frac{{\overline{\delta } {\text{e}}^{{(r + \beta {\text{u}})}} - \underline{\delta } {\text{e}}^{{ - (r + \beta {\text{u}})}} }}{{{\text{e}}^{{(r + \beta {\text{u}})}} + {\text{e}}^{{ - (r + \beta {\text{u}})}} }},$$
(14)

where parameter \({\text{r}}\) is defined as \({\text{r}} = {0}{\text{.5ln}}\left( {\frac{{\underline{\delta } }}{{\overline{\delta } }}} \right)\), and parameter \(\beta\) represents the design parameters. Therefore, the hard saturation function can be expressed as follows:

$$H(u) = S(u) + D(u),$$
(15)

where \(D(u)\) in represents the difference between the hard saturation function \(H(u)\) and soft saturation function \(S(u)\). As shown in Figure 13, when different values of \(\beta\) are taken, the approximate degrees of \(S(u)\) and \(H(u)\) are different. Figure 7 also shows that function \(S(u)\) is a bounded function.

Considering an error between the prediction results of the approximate linear state space of the simulated smoke turbulent flow field and the prediction results of the CFD model of the turbulent flow field, we use variable \(d(t)\) to represent the error between the third-order linear state-space model and the real turbulent flow field. In summary, the linear state-space model of the system becomes the following nonlinear model:

$$\begin{array}{*{20}l} \dot{x}_{1} = x_{2} \hfill \\ \dot{x}_{2} = x_{3} \hfill \\ \dot{x}_{3} = \varphi (x) + H(u) + d(t). \hfill \\ \end{array}$$
(16)
$${\text{The system output is}}\,y = x_{1} .$$
(17)

To realize the equivalence to the real fire smoke flow field based on the closed-loop control method is to realize the target tracking control of the system output \(y\). We define the expected tracking signal of the system as \(y_{d}\), and assume its N + 1 derivative \(y_{d}^{(n)}\), where \(i = 1,2,...,n + 1\), are known and bounded functions of time. Our ultimate goal is to design a control law that can solve the control problem of nonlinear systems under the condition of asymmetric constrained control inputs, and to prove the designed control law theoretically by using Lyapunov stability theory.

We define the residual \(e = x_{1} - y_{d}\) between the simulated smoke concentration at the monitoring position and the desired target. By taking the derivative of the left and right sides of this equation for several times, we can obtain

$$\begin{array}{*{20}l} e = x_{1} - y_{d} \hfill \\ e^{(1)} = \dot{e} = \dot{x}_{1} - \dot{y}_{d} = x_{2} - y_{d}^{(1)} \hfill \\ e^{(2)} = x_{3} - y_{d}^{(2)} \hfill \\ e^{(3)} = \varphi (x) + H(u) + d(t) - y_{d}^{(3)} . \hfill \\ \end{array}$$
(18)

To facilitate the control law design, we define a variable \(s\) to represent the tracking error signal of the nonlinear system (Equation 16). Its mathematical expression is shown as follows:

$$s = \lambda_{1} e + \lambda_{2} e^{(1)} + e^{(2)} ,$$
(19)

where \(\lambda_{1}\), \(\lambda_{2}\) are constants. We can observe that when variable \(s \to 0\), the tracking error signal \(e\), \(e^{(1)}\), \(e^{(2)}\) approaches 0, indicating that the target tracking control of simulated smoke concentration has been realized. Taking the derivative of variables \(s\), we can obtain

$$\begin{aligned} \dot{s} = & \lambda_{1} e^{(1)} + \lambda_{2} e^{(2)} + e^{(3)} \\ = & \varphi (x) + H(u) - y_{d}^{(3)} + \lambda_{1} e^{(1)} + \lambda_{2} e^{(2)} + d(t). \\ \end{aligned}$$
(20)

Let \(\eta = - y_{d}^{(3)} + \lambda_{1} e^{(1)} + \lambda_{2} e^{(2)}\), and the above equation can be written as

$$\begin{aligned} \dot{s} = & \varphi (x) + H(u) + \eta + d(t) \\ = & \varphi (x) + H(u) + \eta + d(t) + c_{1} s - c_{1} s + S(u) - (H(u) - D(u)) \\ = & - c_{1} s + (c_{1} s + \varphi (x) + S(u) + \eta ) + H(u) + (d(t) + D(u)). \\ \end{aligned}$$
(21)

We define variables \(z\) and \(\overline{d}\) as follows:

$$\left\{ \begin{array}{*{20}l} z = c_{1} s + \varphi (x) + S(u) + \eta \hfill \\ \overline{d} = d(t) + D(u) \hfill \\ \end{array} \right.,$$
(22)

where variable \(d(t)\) satisfies the inequality \(d(t) \le d_{\max } < \infty\) and \(d_{\max }\) is the maximum difference between the calculation results of the third-order state space approximation model and the CFD model. Substituting Equation (22) into Equation (21), we can express the variable as follows:

$$\dot{s} = - c_{1} s + z + H(u) + \overline{d}.$$
(23)

Taking the derivative of variables \(z\), we can obtain

$$\dot{z} = c_{1} \dot{s} + \dot{\varphi }(x) + \dot{S}(u) + \dot{\eta }.$$
(24)

The soft saturation function \(S(u)\) is a time-smooth function, so its derivative can be obtained as follows:

$$\dot{S}(u) = \frac{\partial S}{{\partial u}}\dot{u} = g\dot{u}.$$
(25)

The variable \(g\) in Equation (25) can be solved according to the description of \(S(u)\) in Equation (14):

$$\begin{aligned} g = & \frac{{(\overline{\delta } \beta {\text{e}}^{{(r + \beta {\text{u}})}} + \underline{\delta } \beta {\text{e}}^{{ - (r + \beta {\text{u}})}} )({\text{e}}^{{(r + \beta {\text{u}})}} + {\text{e}}^{{ - (r + \beta {\text{u}})}} )}}{{({\text{e}}^{{(r + \beta {\text{u}})}} + {\text{e}}^{{ - (r + \beta {\text{u}})}} )^{2} }} \\ & - \frac{{(\overline{\delta } {\text{e}}^{{(r + \beta {\text{u}})}} - \underline{\delta } {\text{e}}^{{ - (r + \beta {\text{u}})}} )(\beta {\text{e}}^{{(r + \beta {\text{u}})}} - \beta {\text{e}}^{{ - (r + \beta {\text{u}})}} )}}{{({\text{e}}^{{(r + \beta {\text{u}})}} + {\text{e}}^{{ - (r + \beta {\text{u}})}} )^{2} }} \\ = & \frac{{2\beta (\overline{\delta } + \underline{\delta } )}}{{({\text{e}}^{{(r + \beta {\text{u}})}} + {\text{e}}^{{ - (r + \beta {\text{u}})}} )^{2} }}. \\ \end{aligned}$$
(26)

Meanwhile, we let the control signal \(u\) generated by an equivalent control input signal \(v\) get through a first-order filter. Their relationship is shown as follows:

$$\dot{u} = - \alpha u + v,$$
(27)

where \(\alpha\) is a positive constant.

Continuing to process the \(\dot{\varphi }(x)\) in Equation (24), we can obtain

$$\begin{aligned} \dot{\varphi }(x) = & \frac{\partial \varphi }{{\partial x_{3} }}\dot{x}_{3} + \sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} \dot{x}_{i} \\ = & \frac{\partial \varphi }{{\partial x_{3} }}(\varphi (x) + H(u) + d(t)) + \sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} . \\ \end{aligned}$$
(28)

Taking the derivative of \(\eta\) in Equation (24), we can obtain

$$\begin{aligned} \dot{\eta } = & - y_{d}^{(4)} + \lambda_{1} e^{(2)} + \lambda_{2} e^{(3)} \\ = & - y_{d}^{(4)} + \lambda_{1} e^{(2)} + \lambda_{2} (\varphi (x) + H(u) + d(t) - y_{d}^{(3)} ). \\ \end{aligned}$$
(29)
$${\text{We define}}\,\eta_{1} = - y_{d}^{(4)} - \lambda_{2} y_{d}^{(3)} + \lambda_{1} e^{(2)} .$$
(30)

Then, Equation (29) can be changed into

$$\dot{\eta } = \lambda_{2} (\varphi (x) + H(u) + d(t)) + \eta_{1} .$$
(31)

Substituting Equations (25)–(31) into Equation (24), the derivative of variable \(z\) can be further expressed as

$$\begin{aligned} \dot{z} = & c_{1} (\varphi (x) + H(u) + \eta + d(t)) + \sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} + g( - \alpha u + v) \\ & + \frac{\partial \varphi }{{\partial x_{3} }}(\varphi (x) + H(u) + d(t)) + \lambda_{2} (\varphi (x) + H(u) + d(t)) + \eta_{1} \\ = & c_{1} \varphi (x) + c_{1} H(u) + c_{1} \eta + c_{1} d(t) + \sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} - g\alpha u + gv + \frac{\partial \varphi }{{\partial x_{3} }}\varphi (x) \\ & + \frac{\partial \varphi }{{\partial x_{3} }}H(u) + \frac{\partial \varphi }{{\partial x_{3} }}d(t) + \lambda_{2} \varphi (x) + \lambda_{2} H(u) + \lambda_{2} d(t) + \eta_{1} . \\ \end{aligned}$$
(32)

We define function \(A(\cdot)\) as

$$\begin{aligned} A(\cdot) = & c_{1} \varphi (x) + c_{1} H(u) + c_{1} \eta + c_{1} d(t) + \sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} + \frac{\partial \varphi }{{\partial x_{3} }}\varphi (x) \\ & + \frac{\partial \varphi }{{\partial x_{3} }}H(u) + \frac{\partial \varphi }{{\partial x_{3} }}d(t) - g\alpha u + \lambda_{2} \varphi (x) + \lambda_{2} H(u) + \lambda_{2} d(t) + \eta_{1} . \\ \end{aligned}$$
(33)

Substituting Equation (33) into Equation (32), we can obtain the expression of \(\dot{z}\) as follows:

$$\dot{z} = gv + A(\cdot).$$
(34)

We scale the function \(A(\cdot)\) to obtain the following expression:

$$\begin{aligned} A(\cdot) \le & c_{1} \left\| {\varphi (x)} \right\| + c_{1} \Psi + c_{1} \left| \eta \right| + c_{1} d_{\max } + \left| {\frac{\partial \varphi }{{\partial x_{3} }}} \right|\left\| {\varphi (x)} \right\| + \left| {\frac{\partial \varphi }{{\partial x_{3} }}} \right|\Psi \\ & + \left| {\sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} } \right| + \left| {\frac{\partial \varphi }{{\partial x_{3} }}} \right|d_{\max } + \alpha \left| {gu} \right| + \lambda_{2} \left\| {\varphi (x)} \right\| \\ & + \lambda_{2} \Psi + \lambda_{2} d_{\max } + \left| {\eta_{1} } \right| \le aF(\cdot), \\ \end{aligned}$$
(35)

where \(a\) is a positive known constant whose expression is given by Equation (36):

$$a = \max \left\{ {\begin{array}{*{20}c} {c_{1} ,} & {c_{1} \Psi ,} & {c_{1} ,} & {c_{1} d_{\max } ,} & {\Psi ,} & {} \\ {1,} & {d_{\max } ,} & {\alpha ,} & {\lambda_{2} ,} & {\lambda_{2} \Psi ,} & {\lambda_{2} d_{\max } } \\ \end{array} } \right\}.$$
(36)

The definition of the scalar function \(F(\cdot)\) in Equation (35) is shown as

$$F(\cdot) = 4 + 2\left\| {\varphi (x)} \right\| + \left| \eta \right| + \left| {\eta_{1} } \right| + \left| {gu} \right| + \left| {\frac{\partial \varphi }{{\partial x_{3} }}} \right|\left\| {\varphi (x)} \right\| + 2\left| {\frac{\partial \varphi }{{\partial x_{3} }}} \right| + \left| {\sum\limits_{{{\text{i}} = 1}}^{2} {\frac{\partial \varphi }{{\partial x_{i} }}} x_{i + 1} } \right|.$$
(37)

Thus, the inequality can be further written as follows

$$A(\cdot) \le aF(\cdot).$$
(38)

Substituting Equation (38) into Equation (34), we can further obtain the following expression

$$\dot{z} \le gv + aF(\cdot).$$
(39)

For the above nonlinear system with limited control input, the equivalent control input signal \(v\) we designed can be expressed as follows:

$$v = \frac{1}{g}[ - \frac{{azF^{2} }}{\left| z \right|F + \varepsilon } - s - c_{2} z],$$
(40)

where \(\varepsilon\) is an arbitrary small constant and \(c_{2}\) is the control law parameter to be designed.

1.2 Stability Proof

We define a Lyapunov function expression (41) as follows:

$$V = \frac{1}{2}s^{2} + \frac{1}{2}z^{2} .$$
(41)

Taking the derivative of the Lyapunov function \(V(t)\) and substituting Equations (23) and (34), we can obtain

$$\begin{aligned} \dot{V} = & s\dot{s} + z\dot{z} = s( - c_{1} s + z + H(u) + \overline{d}) + z(gv + A) \\ \le & s( - c_{1} s + z + H(u) + \overline{d}) + zgv + \left| z \right|aF. \\ \end{aligned}$$
(42)

According to Equation (40),

$$\begin{aligned} zgv + \left| z \right|aF = & - \frac{{az^{2} F^{2} }}{\left| z \right|F + \varepsilon } - zs - c_{2} z^{2} + \left| z \right|aF \\ = & - c_{2} z^{2} - zs - \frac{{az^{2} F^{2} }}{\left| z \right|F + \varepsilon } + \frac{\left| z \right|aF(\left| z \right|F + \varepsilon )}{{\left| z \right|F + \varepsilon }} \\ \le & - c_{2} z^{2} - zs - \frac{{az^{2} F^{2} }}{\left| z \right|F + \varepsilon } + \frac{{az^{2} F^{2} }}{\left| z \right|F + \varepsilon } + \frac{\left| z \right|Fa\varepsilon }{{\left| z \right|F + \varepsilon }} \\ \le & - c_{2} z^{2} - zs + a\varepsilon . \\ \end{aligned}$$
(43)

Substituting Equation (43) in Equation (42), we can obtain

$$\dot{V} \le - c_{1} s^{2} + sz + sH(u) + s\overline{d} - c_{2} z^{2} - zs + a\varepsilon .$$
(44)

According to Young’s inequality, some terms in the expression of Lyapunov derivative function \(\dot{V}\) are scaled, and the following can be obtained:

$$\begin{gathered} sH(u) \le \frac{1}{2}s^{2} + \frac{1}{2}H^{2} (u) \le \frac{1}{2}s^{2} + \frac{1}{2}\Psi^{2} \hfill \\ s\overline{d} \le \frac{1}{2}s^{2} + \frac{1}{2}\overline{d}^{2} \le \frac{1}{2}s^{2} + \frac{1}{2}d^{2}_{\max } . \hfill \\ \end{gathered}$$
(45)

Substituting Equation (45) into Equation (44), we can obtain

$$\dot{V} \le - c_{1} s^{2} - c_{2} z^{2} + \frac{1}{2}s^{2} + \frac{1}{2}\Psi^{2} + \frac{1}{2}s^{2} + \frac{1}{2}d^{2}_{\max } + a\varepsilon .$$
(46)

We define constant variables \(C = \frac{1}{2}\Psi^{2} + \frac{1}{2}d^{2}_{\max } + a\varepsilon\) and substitute it into Equation (46) to obtain

$$\dot{V} \le - (c_{1} - 1)s^{2} - c_{2} z^{2} + C.$$
(47)

According to Young’s inequality, Equation (47) is further processed, and the following equation can be obtained:

$$\begin{aligned} \dot{V} \le & - (c_{1} - 1)s^{2} - c_{2} z^{2} + C \\ = & - 2(c_{1} - 1)\frac{1}{2}s^{2} - 2c_{2} \frac{1}{2}z^{2} + C \\ \le & - \frac{1}{2}\lambda_{0} s^{2} - \frac{1}{2}\lambda_{0} z^{2} + C \\ = & - \lambda_{0} \left( {\frac{1}{2}s^{2} + \frac{1}{2}z^{2} } \right) + C \\ = & - \lambda_{0} V + C, \\ \end{aligned}$$
(48)

where \(\lambda_{0} = \min \left\{ {2(c_{1} - 1),2c_{2} } \right\}\). The design of parameters \(c_{1}\) in the control law needs to satisfy \(c_{1} - 1 > 0\) to ensure the stability of the closed-loop control system.

Continuing to deal with Equation (48), we multiply both sides of the inequality by \({\text{e}}^{{\lambda_{0} t}}\) and obtain

$$\dot{V}(t){\text{e}}^{{\lambda_{0} t}} \le - \lambda_{0} V(t){\text{e}}^{{\lambda_{0} t}} + C{\text{e}}^{{\lambda_{0} t}} ,$$
(49)

that is,

$$\dot{V}(t){\text{e}}^{{\lambda_{0} t}} + \lambda_{0} V(t){\text{e}}^{{\lambda_{0} t}} \le C{\text{e}}^{{\lambda_{0} t}} .$$
(50)

When Equation (50) is written in differential form, we can obtain

$$\frac{d}{dt}(V(t){\text{e}}^{{\lambda_{0} t}} ) \le C{\text{e}}^{{\lambda_{0} t}} .$$
(51)

We continue to integrate the left and right sides of Equation (51) and obtain

$$V(t){\text{e}}^{{\lambda_{0} t}} \le V(0) + C\int_{0}^{t} {{\text{e}}^{{\lambda_{0} \tau }} } d\tau .$$
(52)

Thus, we can further obtain

$$V(t) \le V(0){\text{e}}^{{ - \lambda_{0} t}} + \frac{C}{{\lambda_{0} }}(1 - {\text{e}}^{{ - \lambda_{0} t}} ) = \frac{C}{{\lambda_{0} }} + \left( {V(0) - \frac{C}{{\lambda_{0} }}} \right){\text{e}}^{{ - \lambda_{0} t}} .$$
(53)

According to the preceding equation, Lyapunov function \(V(t)\) is bounded. When time \(t \to \infty\), the formula for calculating the limit of Lyapunov function \(V(t)\) is as follows

$$\mathop {\lim }\limits_{t \to \infty } V(t) \le \frac{C}{{\lambda_{0} }}.$$
(54)

According to the Lyapunov function expression (41), the following can be obtained:

$$\frac{1}{2}s^{2} \le V(t).$$
(55)

Therefore,

$$\mathop {\frac{1}{2}\lim }\limits_{t \to \infty } s^{2} (t) \le \frac{C}{{\lambda_{0} }}.$$

When time \(t \to \infty\), the following inequality holds:

$$\mathop {\lim }\limits_{t \to \infty } s(t) \le \sqrt {\frac{2C}{{\lambda_{0} }}} .$$
(56)

The preceding analysis shows that the variable \(s\) characterizing the tracking error of the simulated smoke concentration tends to a compact set with the mathematical description \(\Omega : = \left\{ {s:\left| s \right| \le \sqrt {\frac{2C}{{\lambda_{0} }}} } \right\}\). Therefore, the tracking error variables of simulated smoke concentration \(e\) are also globally ultimately bounded. By increasing the value of the parameters \(\lambda_{0}\), set \(\Omega\) can be adjusted to a sufficiently small range. This result means that we can achieve the global uniformly bounded closed-loop control of the tracking error of the simulation smoke concentration by adjusting the control law parameters \(c_{1} ,c_{2}\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, X., Yan, X. & Yang, J. Closed-Loop Control Strategy for Simulated Smoke Concentration in Aircraft Cargo Compartment Mock-Up. Fire Technol 59, 2263–2297 (2023). https://doi.org/10.1007/s10694-023-01433-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10694-023-01433-4

Keywords

Navigation