Skip to main content
Log in

Local controllability of the one-dimensional nonlocal Gray–Scott model with moving controls

  • Published:
Journal of Evolution Equations Aims and scope Submit manuscript

Abstract

In this paper, we prove the local controllability to positive constant trajectories of a nonlinear system of two coupled ODE equations, posed in the one-dimensional spatial setting, with nonlocal spatial nonlinearities, and using only one localized control with a moving support. The model we deal with is derived from the well-known nonlinear reaction–diffusion Gray–Scott model when the diffusion coefficient of the first chemical species \(d_u\) tends to 0 and the diffusion coefficient of the second chemical species \({d_v}\) tends to \(+ \infty \). The strategy of the proof consists in two main steps. First, we establish the local controllability of the reaction–diffusion ODE–PDE derived from the Gray–Scott model taking \(d_u=0\), and uniformly with respect to the diffusion parameter \({d_v} \in (1, +\infty )\). In order to do this, we prove the (uniform) null-controllability of the linearized system thanks to an observability estimate obtained through adapted Carleman estimates for ODE–PDE. To pass to the nonlinear system, we use a precise inverse mapping argument and, secondly, we apply the shadow limit \({d_v} \rightarrow + \infty \) to reduce to the initial system.

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.

Similar content being viewed by others

References

  1. Farid Ammar Khodja, Assia Benabdallah, and Cédric Dupaix. Null-controllability of some reaction-diffusion systems with one control force. J. Math. Anal. Appl., 320(2):928–943, 2006.

  2. Umberto Biccari and Víctor Hernández-Santamaría. Null controllability of linear and semilinear nonlocal heat equations with an additive integral kernel. SIAM J. Control Optim., 57(4):2924–2938, 2019.

    Article  MathSciNet  Google Scholar 

  3. Jean-Michel Coron and Sergio Guerrero. Singular optimal control: a linear 1-D parabolic-hyperbolic example. Asymptot. Anal., 44(3-4):237–257, 2005.

    MathSciNet  MATH  Google Scholar 

  4. Hua Chen, Jian-Meng Li, and Kelei Wang. On the vanishing viscosity limit of a chemotaxis model. Discrete & Continuous Dynamical Systems - A, 40(3):1963–1987, 2020.

    Article  MathSciNet  Google Scholar 

  5. Jean-Michel Coron. Control and nonlinearity, volume 136 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2007.

  6. Felipe W. Chaves-Silva and Mostafa Bendahmane. Uniform null controllability for a degenerating reaction-diffusion system approximating a simplified cardiac model. SIAM J. Control Optim., 53(6):3483–3502, 2015.

    Article  MathSciNet  Google Scholar 

  7. Felipe W. Chaves-Silva and Sergio Guerrero. A uniform controllability result for the Keller-Segel system. Asymptot. Anal., 92(3-4):313–338, 2015.

    MathSciNet  MATH  Google Scholar 

  8. Felipe W. Chaves-Silva, Lionel Rosier, and Enrique Zuazua. Null controllability of a system of viscoelasticity with a moving control. J. Math. Pures Appl. (9), 101(2):198–222, 2014.

  9. Felipe W. Chaves-Silva, Xu Zhang, and Enrique Zuazua. Controllability of evolution equations with memory. SIAM J. Control Optim., 55(4):2437–2459, 2017.

    Article  MathSciNet  Google Scholar 

  10. Arjen Doelman, Tasso J. Kaper, and Paul A. Zegeling. Pattern formation in the one-dimensional Gray-Scott model. Nonlinearity, 10(2):523–563, 1997.

    Article  MathSciNet  Google Scholar 

  11. Asen L. Dontchev. The Graves theorem revisited. J. Convex Anal., 3(1):45–53, 1996.

    MathSciNet  MATH  Google Scholar 

  12. Enrique Fernández-Cara, Juan Límaco, Dany Nina-Huaman, and Miguel R. Núñez Chávez. Exact controllability to the trajectories for parabolic PDEs with nonlocal nonlinearities. Math. Control Signals Systems, 31(3):415–431, 2019.

  13. Enrique Fernández-Cara, Qi Lü, and Enrique Zuazua. Null controllability of linear heat and wave equations with nonlocal spatial terms. SIAM J. Control Optim., 54(4):2009–2019, 2016.

    Article  MathSciNet  Google Scholar 

  14. Andrei V. Fursikov and Oleg Yu. Imanuvilov. Controllability of evolution equations, volume 34 of Lecture Notes Series. Seoul National University, Research Institute of Mathematics, Global Analysis Research Center, Seoul, 1996.

  15. Sergio Guerrero and Gilles Lebeau. Singular optimal control for a transport-diffusion equation. Comm. Partial Differential Equations, 32(10-12):1813–1836, 2007.

    Article  MathSciNet  Google Scholar 

  16. Patricio Guzman and Lionel Rosier. Null Controllability of the Structurally Damped Wave Equation on the Two-Dimensional Torus. SIAM J. Control Optim., 59(1):131–155, 2021.

    Article  MathSciNet  Google Scholar 

  17. Mamadou Gueye. Insensitizing controls for the Navier-Stokes equations. Ann. Inst. H. Poincaré Anal. Non Linéaire, 30(5):825–844, 2013.

  18. Víctor Hernández-Santamaría and Kévin Le Balc’h. Local Null-Controllability of a Nonlocal Semilinear Heat Equation. Applied Mathematics & Optimization, 2020.

  19. Víctor Hernández-Santamaría and Enrique Zuazua. Controllability of shadow reaction-diffusion systems. J. Differential Equations, 268(7):3781–3818, 2020.

    Article  MathSciNet  Google Scholar 

  20. Karl Kunisch and Diego A. Souza. On the one-dimensional nonlinear monodomain equations with moving controls. J. Math. Pures Appl. (9), 117:94–122, 2018.

    Article  MathSciNet  Google Scholar 

  21. Pierre Lissy and Enrique Zuazua. Internal controllability for parabolic systems involving analytic non-local terms. Chin. Ann. Math. Ser. B, 39(2):281–296, 2018.

    Article  MathSciNet  Google Scholar 

  22. Anna Marciniak-Czochra, Steffen Härting, Grzegorz Karch, and Kanako Suzuki. Dynamical spike solutions in a nonlocal model of pattern formation. Nonlinearity, 31(5):1757–1781, 2018.

    Article  MathSciNet  Google Scholar 

  23. Jeff S. McGough and Kyle Riley. Pattern formation in the Gray-Scott model. Nonlinear Anal. Real World Appl., 5(1):105–121, 2004.

    Article  MathSciNet  Google Scholar 

  24. Philippe Martin, Lionel Rosier, and Pierre Rouchon. Null controllability of the structurally damped wave equation with moving control. SIAM J. Control Optim., 51(1):660–684, 2013.

    Article  MathSciNet  Google Scholar 

  25. Michel Pierre. Global existence in reaction-diffusion systems with control of mass: a survey. Milan J. Math., 78(2):417–455, 2010.

    Article  MathSciNet  Google Scholar 

  26. Jacques Simon. Compact sets in the space \(L^p(0,T;B)\). Ann. Mat. Pura Appl. (4), 146:65–96, 1987.

    MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank the anonymous referees for their careful reading and valuable comments and suggestions that helped us to improve the presentation of this paper. This work has received support of the FONDECYT project No 3200028 of ANID, Chile. This study has been carried out with financial support from the French State, managed by the French National Research Agency (ANR) in the frame of the “Investments for the future” Programme IdEx Bordeaux - SysNum (ANR-10-IDEX-03-02). This work has also been supported by the grant “Numerical simulation and optimal control in view of temperature regulation in smart buildings” of the Nouvelle Aquitaine Region. The first author has also been partially supported by the program “Estancias posdoctorales por México” of CONACyT, Mexico.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kévin Le Balc’h.

Additional information

Publisher's Note

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

Appendices

A Uniform parabolic Carleman estimate

In this part, we prove Lemma 2.3. The proof follows the arguments presented in [20, Appendix A] with some remarks coming grom [6]. For completeness, we give a sketch of the proof.

We start by giving some useful properties on the weight function \(\alpha \) and its derivatives. By simple computations using (30), we have

$$\begin{aligned} \begin{aligned} \alpha _x&=-\lambda \xi \eta _x, \\ \alpha _{xx}&=\lambda ^2\xi (-\eta _x^2-\lambda ^{-1}\eta _{xx}), \\ |\alpha _{xx}|&\le C\lambda ^2\xi , \\ |\alpha _{xxx}|&\le C\lambda ^3\xi , \\ |\alpha _{xxxx}|&\le C\lambda ^4\xi , \\ |\alpha _t|&\le C(T+e^{2\lambda \Vert \eta \Vert _\infty })\lambda \xi ^2, \\ |\alpha _{xt}|&\le C(T+1)\lambda ^2\xi ^2, \\ |\alpha _{tt}|&\le C(T+T^2+e^{2\lambda \Vert \eta \Vert _\infty })\lambda ^2\xi ^3. \end{aligned} \end{aligned}$$
(90)

As usual, we set \(w=e^{-s\alpha }\psi \) and compute

$$\begin{aligned} w_{x}=-s\alpha _x w+e^{-s\alpha }\psi _x. \end{aligned}$$

Using the boundary conditions of \(\psi \), we deduce \(w_x=-s\alpha _xw\) on \((0,T)\times \{0,1\}\). We introduce the parabolic operator \(P=\partial _t+\frac{1}{\epsilon }\partial _{xx}\). Then, we have

$$\begin{aligned} e^{-s\alpha }P(e^{s\alpha }w)=P_e^{\epsilon }w+P_{k}^{\epsilon }w, \end{aligned}$$
(91)

where

$$\begin{aligned} P_e^{\epsilon }w&=\frac{1}{\epsilon } w_{xx}+s\alpha _t w+\frac{1}{\epsilon }s^2\alpha _x^2 w, \end{aligned}$$
(92)
$$\begin{aligned} P_k^{\epsilon }w&=w_t+\frac{2}{\epsilon }s\alpha _xw_x+\frac{1}{\epsilon }s\alpha _{xx}w. \end{aligned}$$
(93)

We take the \(L^2\)-norm in both sides of (91), thus obtaining

$$\begin{aligned}&\!\!\!\Vert e^{-s\alpha } P(e^{s\alpha }w)\Vert _{L^2(Q_T)}^2=\Vert P_{e}^\epsilon w\Vert ^2_{L^2(Q_T)} +\Vert P_k^\epsilon w\Vert ^2_{L^2(Q_T)}+2\left( P_e^\epsilon w,P_k^\epsilon w\right) _{L^2(Q_T)}.\nonumber \\ \end{aligned}$$
(94)

A very long, but straightforward computation gives that

$$\begin{aligned} 2\left( P_e^\epsilon w,P_k^\epsilon w\right) _{L^2(Q_T)}= BT+DT_1+DT_2, \end{aligned}$$

where

$$\begin{aligned} BT:=&\frac{1}{\epsilon }\left. \int _{0}^{T}s\alpha _{xt}w^2\text {d}t\right| _{0}^{1}+\frac{2}{\epsilon ^2}\left. \int _{0}^{T}s\alpha _{x}w_x^2\text {d}t\right| _{0}^{1}+\frac{2}{\epsilon ^2}\left. \int _{0}^{T}s\alpha _{xx}w_x w\text {d}t\right| _{0}^{1} \\&-\frac{1}{\epsilon ^2}\left. \int _{0}^{T}s\alpha _{xxx}w^2\text {d}t\right| _{0}^{1}+\frac{2}{\epsilon }\left. \int _{0}^{T}s^2\alpha _t\alpha _{x}w^2\text {d}t\right| _{0}^{1}+\frac{2}{\epsilon ^2}\left. \int _{0}^{T}s^3\alpha _{x}^3 w^2\text {d}t\right| _{0}^{1}, \\ DT_1:=&-\frac{4}{\epsilon ^2}\iint _{Q_T} s\alpha _{xx}w_x^2\,\text {d}x\text {d}t, \\ DT_2:=&-\frac{4}{\epsilon }\iint _{Q_T}s^2\alpha _{tx}\alpha _x w^2\,\text {d}x\text {d}t+\frac{1}{\epsilon ^2}\iint _{Q_T}s\alpha _{xxxx}w^2\,\text {d}x\text {d}t-\iint _{Q_T}s\alpha _{tt}w^2\,\text {d}x\text {d}t\\&-\frac{4}{\epsilon ^2}\iint _{Q_T} s^3\alpha _x^2\alpha _{xx}w^2\,\text {d}x\text {d}t. \end{aligned}$$

Using that \(w_{x}=-s\alpha _x w\) in \((0,T)\times \{0,1\}\), we can obtain

$$\begin{aligned} BT=&\frac{4}{\epsilon ^2}\left. \int _{0}^T s^3\alpha _x^3 w^2 \text {d}t\right| _{0}^{1}+\frac{2}{\epsilon }\left. \int _{0}^{T}s^2\alpha _t\alpha _{x}w^2\text {d}t\right| _{0}^{1}\frac{2}{\epsilon ^2}\left. \int _{0}^{T}s^2\alpha _x\alpha _{xx}w^2\text {d}t\right| _{0}^{1} \\&+\frac{1}{\epsilon }\left. \int _{0}^{T}s\alpha _{xt}w^2\text {d}t\right| _{0}^{1}-\frac{1}{\epsilon ^2}\left. \int _{0}^{T}s\alpha _{xxx}w^2\text {d}t\right| _{0}^{1}. \end{aligned}$$

Moreover, using properties (90) together with (25)–(26), it is not difficult to see that

$$\begin{aligned} BT\ge&\frac{4C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text{ d }t\Big |_{x=1}+\frac{4C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text{ d }t\Big |_{x=0} \\ {}&-\frac{2C}{\epsilon ^2}\int _{0}^{T}s^2\lambda ^2\xi ^3(T+e^{2\lambda \Vert \eta \Vert _\infty })w^2\text{ d }t\Big |_{x=1}\\ {}&-\frac{2C}{\epsilon ^2}\int _{0}^{T}s^2\lambda ^2\xi ^3(T+e^{2\lambda \Vert \eta \Vert _\infty })w^2\text{ d }t\Big |_{x=0} \\ {}&-\frac{2}{\epsilon ^2}\int _{0}^{T}s^2\lambda ^3\xi ^2w^2\text{ d }t\Big |_{x=1}-\frac{2}{\epsilon ^2}\int _{0}^{T}s^2\lambda ^3\xi ^2w^2\text{ d }t\Big |_{x=0} \\ {}&-\frac{C}{\epsilon ^2}\int _{0}^{T}s(T+1)\lambda ^2\xi ^2w^2\text{ d }t\Big |_{x=1}-\frac{C}{\epsilon ^2}\int _{0}^{T}s(T+1)\lambda ^2\xi ^2w^2\text{ d }t\Big |_{x=0} \\ {}&-\frac{C}{\epsilon ^2}\int _{0}^{T}s\lambda ^3\xi w^2\text{ d }t\Big |_{x=1}-\frac{C}{\epsilon ^2}\int _{0}^{T}s\lambda ^3\xi w^2\text{ d }t\Big |_{x=0}, \end{aligned}$$

where we have used that \(0<\epsilon \le 1\) to adjust the powers of \(\epsilon \). Finally, taking \(\lambda \ge C\) and \(s\ge C(1+T+e^{2\lambda \Vert \eta \Vert _\infty })\), we get

$$\begin{aligned}&BT\ge \frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=1} +\frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=0}. \end{aligned}$$
(95)

Let us focus now on the distributed terms. Using (90), we can rewrite

$$\begin{aligned}&DT_1=\frac{4}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi \eta _x^2w_x^2\,\text{ d }x\text{ d }t+\frac{4}{\epsilon ^2}\iint _{Q_T}s\lambda \xi \eta _{xx}w_x^2\,\text{ d }x\text{ d }t, \end{aligned}$$

and

$$\begin{aligned} DT_2=&-\frac{4}{\epsilon }\iint _{Q_T}s^2\alpha _{xt}\alpha _x w^2\,\text {d}x\text {d}t+\frac{1}{\epsilon ^2}\iint _{Q_T}s\alpha _{xxxx} w^2\,\text {d}x\text {d}t\\&-\iint _{Q_T}s\alpha _{tt}w^2\,\text {d}x\text {d}t+\frac{4}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^3\xi ^3\eta _x^2\eta _{xx}w^2\,\text {d}x\text {d}t+\frac{4}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3\eta _x^4w^2\,\text {d}x\text {d}t. \end{aligned}$$

Using estimates (90), we can bound by below as follows

$$\begin{aligned}&DT_1\ge \frac{4}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi \eta _x^2w_x^2\,\text {d}x\text {d}t-\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda \xi w_x^2\,\text {d}x\text {d}t. \end{aligned}$$
(96)

and

$$\begin{aligned} DT_2\ge&\frac{4}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3\eta _x^4w^2\,\text {d}x\text {d}t-\frac{C(T+1)}{\epsilon }\iint _{Q_T}s^2\lambda ^3\xi ^3 w^2\,\text {d}x\text {d}t-\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda ^4\xi w^2\,\text {d}x\text {d}t\nonumber \\&-C\left( T+T^2+e^{2\lambda \Vert \eta \Vert _\infty }\right) \iint _{Q_T}s\lambda ^2\xi ^3w^2\,\text {d}x\text {d}t-\frac{C}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^3\xi ^3w^2\,\text {d}x\text {d}t. \end{aligned}$$
(97)

Collecting estimates (95)–(97) yield

$$\begin{aligned} 2\left( P_e^\epsilon w,P_k^\epsilon w\right) _{L^2(Q_T)} \ge&\frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=1} +\frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=0} \\&+\frac{4}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi \eta _x^2w_x^2\,\text {d}x\text {d}t+\frac{4}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3\eta _x^4w^2\,\text {d}x\text {d}t\\&-\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda \xi w_x^2\,\text {d}x\text {d}t-\frac{C(T+1)}{\epsilon }\iint _{Q_T}s^2\lambda ^3\xi ^3 w^2\,\text {d}x\text {d}t\\&-\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda ^4\xi w^2\,\text {d}x\text {d}t-C\left( T+T^2+e^{2\lambda \Vert \eta \Vert _\infty }\right) \\&\times \iint _{Q_T}s\lambda ^2\xi ^3w^2\,\text {d}x\text {d}t-\frac{C}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^3\xi ^3w^2\,\text {d}x\text {d}t. \end{aligned}$$

Using property (21) and taking \(\lambda \ge C\) and \(s\ge C(T+T^2)\), we deduce

$$\begin{aligned}&2\left( P_e^\epsilon w,P_k^\epsilon w\right) _{L^2(Q_T)} +\frac{1}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s^3\lambda ^4\xi ^3 w^2 \,\text {d}x\text {d}t+\frac{1}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s\lambda ^2\xi w_x^2 \,\text {d}x\text {d}t\nonumber \\&\quad \ge \frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=1}+\frac{C}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=0} +\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi w_x^2\,\text {d}x\text {d}t\nonumber \\&\qquad +\frac{C}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3 w^2\,\text {d}x\text {d}t. \end{aligned}$$
(98)

Combining (98) with (94), we get

$$\begin{aligned} \begin{aligned}&C\Vert e^{-s\alpha } P(e^{s\alpha }w)\Vert _{L^2(Q_T)}^2 +\frac{C}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s^3\lambda ^4\xi ^3 w^2 \,\text {d}x\text {d}t\\&\qquad +\frac{C}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s\lambda ^2\xi w_x^2 \,\text {d}x\text {d}t\\&\quad \ge \Vert P_{e}^\epsilon w\Vert ^2_{L^2(Q_T)} +\Vert P_k^\epsilon w\Vert ^2_{L^2(Q_T)}+\frac{1}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=1}\\&\qquad +\frac{1}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=0} \\&\qquad +\frac{1}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi w_x^2\,\text {d}x\text {d}t+\frac{1}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3 w^2\,\text {d}x\text {d}t. \end{aligned} \end{aligned}$$
(99)

We will add terms corresponding to \(w_{xx}\) and \(w_t\) to the right-hand side of (99). For this, we multiply (92) by \(s^{-1/2}\xi ^{-1/2}\) and take the \(L^2\)-norm, that is,

$$\begin{aligned} \frac{1}{\epsilon ^2}s^{-1}\Vert \xi ^{-1/2}w_{xx}\Vert ^2_{L^2(Q_T)}&=s^{-1}\Vert \xi ^{-1/2}(P_e^{\epsilon }w-s\alpha _tw\nonumber \\ {}&\quad -\frac{1}{\epsilon }s^2\alpha _x^2 w)\Vert ^2_{L^2(Q_T)} \nonumber \le C s^{-1}\iint _{Q_T}\xi ^{-1}|P_e^{\epsilon }w|^2\,\text{ d }x\text{ d }t\nonumber \\ {}&\quad +CT^2\iint _{Q_T}s^2\lambda ^2\xi ^3w^2\,\text{ d }x\text{ d }t\nonumber +\frac{C}{\epsilon ^2}s^3\iint _{Q_T}\lambda ^4\xi ^3 w^2\,\text{ d }x\text{ d }t\nonumber \\ {}&\le C s^{-1}\!\iint _{Q_T}\xi ^{-1}|P_e^{\epsilon }w|^2\,\text{ d }x\text{ d }t\!+\!\frac{C}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3w^2\,\text{ d }x\text{ d }t, \end{aligned}$$
(100)

where we have used that \(\epsilon \le 1\) and \(s\ge CT^2\) in the last line. Arguing in the same way and considering (93), it is not difficult to see that

$$\begin{aligned} s^{-1}\Vert \xi ^{-1/2}w_t\Vert ^2_{L^2(Q_T)}\nonumber&\le Cs^{-1}\iint _{Q_T}\xi ^{-1}|P_k^{\epsilon }w|^2\,\text {d}x\text {d}t\nonumber \\&\quad +\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi w_{x}^2\,\text {d}x\text {d}t +\frac{C}{\epsilon ^2}\iint _{Q_T}s\lambda ^4\xi w^2\,\text {d}x\text {d}t. \end{aligned}$$
(101)

Using that \(s^{-1}\xi ^{-1}\le C\) for all \((t,x)\in Q_T\), we can use (100) and (101) to estimate from below in (99) and obtain

$$\begin{aligned}&C\Vert e^{-s\alpha } P(e^{s\alpha }w)\Vert _{L^2(Q_T)}^2\nonumber +\frac{C}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s^3\lambda ^4\xi ^3 w^2 \,\text {d}x\text {d}t\nonumber \\&\qquad +\frac{C}{\epsilon ^2}\iint _{\omega _{0}(t)\times (0,T)}s\lambda ^2\xi w_x^2 \,\text {d}x\text {d}t\nonumber \\&\quad \ge \ \frac{1}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=1} +\frac{1}{\epsilon ^2}\int _0^Ts^3\lambda ^3\xi ^3w^2\text {d}t\Big |_{x=0}\nonumber \\&\qquad +\frac{1}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi w_x^2\,\text {d}x\text {d}t +\frac{1}{\epsilon ^2}\iint _{Q_T}s^3\lambda ^4\xi ^3w^2\,\text {d}x\text {d}t\nonumber \\&\qquad + \frac{1}{\epsilon ^2}\iint _{Q_T}s^{-1}\xi ^{-1}|w_{xx}|^2\,\text {d}x\text {d}t+\iint _{Q_T}s^{-1}\xi ^{-1}|w_{t}|^2\,\text {d}x\text {d}t. \end{aligned}$$
(102)

To conclude the proof, we need to eliminate the local term of \(w_x\) in the above equation. For this, consider a function \(\zeta \in C^\infty ([0,T]\times [0,L])\) verifying

$$\begin{aligned} {\left\{ \begin{array}{ll} 0\le \zeta \le 1 &{}\forall (t,x)\in [0,T]\times [0,L], \\ \zeta (t,x)=1 &{}\forall t\in [0,T], \;\; \forall x\in \omega _0(t),\\ \zeta (t,x)=0 &{}\forall t\in [0,T], \;\; \forall x\in [0,L]{\setminus }\overline{\omega _1(t)}. \end{array}\right. } \end{aligned}$$

We have

$$\begin{aligned} \frac{1}{\epsilon ^2}\iint _{\omega _0(t)\times (0,T)}s\lambda ^2\xi w_{x}^2\,\text {d}x\text {d}t&\le \frac{1}{\epsilon ^2}\iint _{Q_T}\zeta s\lambda ^2\xi w_{x}^2\,\text {d}x\text {d}t\\&=-\frac{1}{\epsilon ^2}\int _{0}^{T}s\lambda ^2\xi \zeta w w_{x}\text {d}t\Big |_{0}^{1} -\frac{1}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi ww_{xx}\zeta \,\text {d}x\text {d}t\\&\quad -\frac{1}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi _{x} ww_{x}\zeta \,\text {d}x\text {d}t-\frac{1}{\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi ww_{x}\zeta _{x}\,\text {d}x\text {d}t. \end{aligned}$$

Using the fact that \(w_x=-s\alpha _x w\) on \((0,T)\times \{0,1\}\) together with properties (25)–(26) and Cauchy–Schwarz and Young inequalities, we get

$$\begin{aligned} \frac{C}{\epsilon ^2}\iint _{w_0(t)\times (0,T)}s\lambda ^2\xi w_x^2&\le \ \frac{1}{2\epsilon ^2}\iint _{Q_T}s^{-1}\xi ^{-1}|w_{xx}|^2\,\text {d}x\text {d}t\\&\quad +\frac{1}{2\epsilon ^2}\iint _{Q_T}s\lambda ^2\xi w_x^2\,\text {d}x\text {d}t+\frac{C}{\epsilon ^2}\iint _{\omega _1(t)\times (0,T)}s^3\lambda ^4\xi ^3w^2\,\text {d}x\text {d}t. \end{aligned}$$

Using the above estimate in (102) and recalling the change of variables \(w=e^{-s\alpha }\psi \) gives the desired result. This concludes the proof.

Proof of a precise observability inequality from the Carleman estimate

The goal of this part is to prove Proposition 2.6.

Proof

The proof is by now standard and relies on well-known arguments, we follow here the presentation in [17, Lemma 4.1]. We fix \(\lambda , s\) large enough such that the Carleman estimate from Proposition 2.5 holds and such that we have the following estimate

$$\begin{aligned} e^{-4s\alpha ^\star +2s\widehat{\alpha }}\le C e^{-s\alpha }\le C e^{-s\alpha ^\star } . \end{aligned}$$
(103)

All the following positive constants \(C>0\) can now depend on \(\lambda ,s\).

By construction, \(\alpha =\beta \) and \(\xi =\gamma \) in \((T/2,T)\times (0,1)\), therefore

$$\begin{aligned}&\int _{T/2}^{T} \int _0^1 \gamma |\phi |^2 e^{-2s \beta } \,\text{ d }x\text{ d }t\nonumber +\int _{T/2}^{T} \int _0^1 \gamma ^3 |\psi |^2 e^{-2s \beta } \,\text{ d }x\text{ d }t\nonumber \\ {}&\quad = \int _{T/2}^{T} \int _0^1 \xi |\phi |^2 e^{-2s \alpha } \,\text{ d }x\text{ d }t+\int _{T/2}^{T} \int _0^1 \xi ^3 |\psi |^2 e^{-2s \alpha } \,\text{ d }x\text{ d }t. \end{aligned}$$
(104)

Moreover, from the definition of \(\beta ^\star \) and \(\alpha ^\star \), we readily see that \(e^{-s\alpha ^\star }\le e^{-s\beta ^\star }\).

From this fact and noting that \(\beta \) (resp. \(\alpha \)) blows up exponentially as \(t\rightarrow T^{-}\) (resp. \(t\rightarrow 0^+\) and \(t\rightarrow T^{-}\)) while \(\gamma \) (resp. \(\xi \)) blows up polynomially as \(t\rightarrow T^{-}\) (resp. \(t\rightarrow 0^+\) and \(t\rightarrow T^{-}\)), we can use (104), (103) and our Carleman inequality (34) to deduce that

$$\begin{aligned}&\int _{T/2}^{T} \int _0^1 \gamma |\phi |^2 e^{-2s \beta } \,\text {d}x\text {d}t\nonumber + \int _{T/2}^{T} \int _0^1 \gamma ^3 |\psi |^2 e^{-2s \beta } \,\text {d}x\text {d}t\nonumber \\&\quad \le \ C(s,\lambda ) \Bigg ( \iint _{Q_T}e^{-2s\beta }\gamma ^3|g_1|^2\,\text {d}x\text {d}t+ \iint _{Q_T}e^{-2s\beta } |g_2|^2\,\text {d}x\text {d}t\nonumber \\&\qquad + \iint _{\omega _2\times (0,T)} e^{-s\beta ^\star }({\gamma }^\star )^{8}|\phi |^2\,\text {d}x\text {d}t\Bigg ) . \end{aligned}$$
(105)

For the set \((0,T/2)\times (0,1)\), we will use energy estimates for the system (18). More precisely, consider the function \(\nu \in C^1([0,T])\) such that

$$\begin{aligned} \nu =1 \text { in } [0,T/2], \quad \nu =0 \text { in }[3T/4,T/4], \quad |\nu ^\prime (t)|\le C/T. \end{aligned}$$
(106)

Setting \(({\widetilde{\phi }},{\widetilde{\psi }})=(\nu \phi ,\nu \psi )\), it is not difficult to see that the new variables verify the system

$$\begin{aligned} {\left\{ \begin{array}{ll} -\widetilde{\phi }_t=a_{11} {\widetilde{\phi }} + a_{21}{\widetilde{\psi }} + \nu g_1-\nu ^\prime \phi &{} \text {in } Q_T, \\ -{\widetilde{\psi }}_t={d_v}{\widetilde{\psi }}_{xx} +a_{12}{\widetilde{\phi }}+a_{22} {\widetilde{\psi }} +\nu g_2 - \nu ^\prime \psi &{}\text {in } Q_T, \\ \partial _x {\widetilde{\psi }}=0 &{}\text {on } \Sigma _T, \\ ({\widetilde{\phi }},{\widetilde{\psi }})(T,\cdot )=(0,0) &{}\text {in }(0,1). \end{array}\right. } \end{aligned}$$
(107)

From standard energy estimates, we deduce that system (107) verifies

$$\begin{aligned}&\int _{\Omega }|{\widetilde{\phi }}(t,x)|^2\,\text {d}x+\int _{\Omega }|\widetilde{\psi }(t,x)|^2\,\text {d}x+{d_v} \int _{t}^{T}\!\!\!\int _{\Omega }|\widetilde{\psi }_x|^2\,\text {d}x\text {d}t\\&\quad \le \ C\left( \int _{t}^{T}\!\!\!\int _{\Omega }|{\widetilde{\phi }}|^2\,\text {d}x\text {d}t\right. +\int _{t}^{T}\!\!\!\int _{\Omega }|{\widetilde{\psi }}|^2\,\text {d}x\text {d}t+\int _{t}^{T}\!\!\!\int _{\Omega }|\eta g_1|^2\,\text {d}x\text {d}t\\&\qquad +\int _{t}^{T}\!\!\!\int _{\Omega }|\eta g_2|^2\,\text {d}x\text {d}t+ \int _{t}^{T}\!\!\!\int _{\Omega }|\eta ^\prime \phi |^2\,\text {d}x\text {d}t\left. +\int _{t}^{T}\!\!\!\int _{\Omega }|\eta ^\prime \psi |^2\,\text {d}x\text {d}t\right) , \quad \forall t\in [0,T], \end{aligned}$$

where C is a positive constant only depending on \(a_{ij}\).

Dropping the third term in the left-hand side of the above expression, we use Gronwall’s inequality to deduce

$$\begin{aligned}&\Vert {\widetilde{\phi }}(0, \cdot )\Vert ^2_{L^2(\Omega )} + \Vert {\widetilde{\psi }}(0, \cdot )\Vert ^2_{L^2(\Omega )} + \iint _{Q_T}|{\widetilde{\phi }}|^2 \,\text{ d }x\text{ d }t+ \iint _{Q_T}|{\widetilde{\psi }}|^2\,\text{ d }x\text{ d }t\\ {}&\quad \le C\left( \iint _{Q_T} |\eta g_1|^2\,\text{ d }x\text{ d }t+\iint _{Q_T}|\eta g_2|^2\,\text{ d }x\text{ d }t\right. \left. + \iint _{Q_T}|\eta ^\prime \phi |^2 \,\text{ d }x\text{ d }t\right. \nonumber \\ {}&\quad \left. +\iint _{Q_T}|\eta ^\prime \psi |^2\,\text{ d }x\text{ d }t\right) , \end{aligned}$$

for some constant \(C>0\) only depending on T and \(a_{ij}\).

Recalling the definition of \(\eta \), we obtain from the above expression

$$\begin{aligned}&\Vert \phi (0, \cdot )\Vert ^2_{L^2(\Omega )} + \Vert \psi (0, \cdot )\Vert ^2_{L^2(\Omega )} + \int _{0}^{T/2}\!\!\!\!\int _{\Omega } | \phi |^2 \,\text {d}x\text {d}t+ \int _{0}^{T/2}\!\!\!\!\int _{\Omega }| \psi |^2\,\text {d}x\text {d}t\\&\quad \le C\left( \int _{0}^{3T/4}\!\!\!\!\int _{\Omega }| g_1|^2 \,\text {d}x\text {d}t\right. \left. +\int _{0}^{3T/4}\!\!\!\!\int _{\Omega }| g_2|^2\,\text {d}x\text {d}t\right) + \frac{C}{T^2}\left( \int _{T/2}^{3T/4}\!\!\!\!\int _{\Omega } |\phi |^2 \,\text {d}x\text {d}t\right. \\&\qquad \left. +\int _{T/2}^{3T/4}\!\!\!\!\int _{\Omega }| \psi |^2\,\text {d}x\text {d}t\right) . \end{aligned}$$

Since the domain of integration in the above integrals is away from the singularity of the weight functions (48) at \(t=T\) (and therefore they are bounded), we can introduce them in the above inequality as follows

$$\begin{aligned}&\Vert \phi (0, \cdot )\Vert ^2_{L^2(\Omega )} + \Vert \psi (0, \cdot )\Vert ^2_{L^2(\Omega )} + \int _{0}^{T/2}\!\!\!\!\int _{\Omega } e^{-2s\beta } \gamma \left( | \phi |^2 + | \phi _x|^2 \right) \,\text {d}x\text {d}t\\&\qquad + \int _{0}^{T/2}\!\!\!\!\int _{\Omega } \gamma ^3 | \psi |^2 e^{-2s\beta }\,\text {d}x\text {d}t\\&\quad \le C(s,\lambda ,T)\left( \int _{0}^{3T/4}\!\!\!\!\int _{\Omega } e^{-2s\beta } \gamma ^3 | g_1|^2 \,\text {d}x\text {d}t\right. \left. +\int _{0}^{3T/4}\!\!\!\!\int _{\Omega }e^{-2s\beta } | g_2|^2\,\text {d}x\text {d}t\right) \\&\qquad + C(s,\lambda ,T)\left( \int _{T/2}^{3T/4}\!\!\!\!\int _{\Omega } e^{-2s\beta }\gamma |\phi |^2 \,\text {d}x\text {d}t\right. \left. +\int _{T/2}^{3T/4}\!\!\!\!\int _{\Omega } e^{-2s\beta }\gamma ^3| \psi _2|^2\,\text {d}x\text {d}t\right) . \end{aligned}$$

Using estimate (105) to bound all the terms on the last line of the above inequality and adding up the resulting expression to (105) yields

$$\begin{aligned}&\Vert \phi (0, \cdot )\Vert ^2_{L^2(\Omega )} +\Vert \psi (0, \cdot )\Vert ^2_{L^2(\Omega )} + \iint _{Q_T} e^{-2s\beta } \gamma | \phi |^2 \,\text {d}x\text {d}t+ \iint _{Q_T} \gamma ^3 | \psi |^2 e^{-2s\beta }\,\text {d}x\text {d}t\\&\quad \le C\left( \iint _{Q_T} e^{-2s\beta } \gamma ^3 | g_1|^2 \,\text {d}x\text {d}t+\iint _{Q_T} e^{-2s\beta } | g_2|^2\,\text {d}x\text {d}t\right) \\&\qquad +C\left( \iint _{\omega _2(t)\times (0,T)} e^{-s\beta ^\star }({\gamma }^\star )^{8}|\phi |^2 \,\text {d}x\text {d}t\right) . \end{aligned}$$

To conclude, it is enough to use definitions (48) in the above inequality. This ends the proof. \(\square \)

Some properties of the heat semigroup

We recall in the next result some well-known facts about the heat semigroup with a diffusion parameter \(d_v >0 \) with homogeneous Neumann boundary conditions on the interval (0, 1), denoted by \(\{e^{t{d_v}\partial _{xx}}\}_{t\ge 0}\). The proof can be found for instance in [22, Lemma A.1].

Lemma C.1

The following properties hold true.

  1. a.

    For every constant \(K\in {\mathbb {R}}\), we have \(e^{t{d_v} \partial _{xx}}K=K\) for all \(t\ge 0\).

  2. b.

    For every \(z_0\in L^2(\Omega )\), there exists a constant \(C>0\) only depending on \(z_0\) such that for every \(t \ge 0\),

    $$\begin{aligned} \left\Vert e^{t{d_v}\partial _{xx}}\left( z_0-\int _{\Omega }z_0\,\text {d}x\right) \right\Vert _{L^2(\Omega )}\le C e^{-\lambda _1 d_v t} \left\Vert z_0\right\Vert _{L^2(\Omega )}, \end{aligned}$$
    (108)

    where \(\lambda _1 >0\) is the first positive eigenvalue of the Neumann Laplacian operator \(-\partial _{xx}\) on (0, 1).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hernández-Santamaría, V., Le Balc’h, K. Local controllability of the one-dimensional nonlocal Gray–Scott model with moving controls. J. Evol. Equ. 21, 4539–4574 (2021). https://doi.org/10.1007/s00028-021-00725-y

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00028-021-00725-y

Keywords

Mathematics Subject Classification

Navigation