Skip to main content
Log in

Existence of Global-in-Time Solutions to a System of Fully Nonlinear Parabolic Equations

  • Published:
Acta Applicandae Mathematicae Aims and scope Submit manuscript

Abstract

We consider the Cauchy problem for a system of fully nonlinear parabolic equations. In this paper, we shall show the existence of global-in-time solutions to the problem. Our condition to ensure the global existence is specific to the fully nonlinear parabolic 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

Data Availability

Not applicable.

Materials Availability

Not applicable.

Code Availability

Not applicable.

References

  1. Armstrong, S.N., Trokhimtchouk, M.: Long-time asymptotics for fully nonlinear homogeneous parabolic equations. Calc. Var. Partial Differ. Equ. 38, 521–540 (2010)

    Article  MathSciNet  Google Scholar 

  2. Crandall, M.G., Ishii, H., Lions, P.-L.: User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. 27, 1–67 (1992)

    Article  MathSciNet  Google Scholar 

  3. Crandall, M.G., Lions, P.-L.: Quadratic growth of solutions of fully nonlinear second order equations in \(\mathbb{R}^{n}\). Differ. Integral Equ. 3, 601–616 (1990)

    MATH  Google Scholar 

  4. Deng, K., Levine, H.A.: The role of critical exponents in blow-up theorems: the sequel. J. Math. Anal. Appl. 243, 85–126 (2000)

    Article  MathSciNet  Google Scholar 

  5. Escobedo, M., Herrero, M.A.: Boundedness and blow up for a semilinear reaction-diffusion system. J. Differ. Equ. 89, 176–202 (1991)

    Article  MathSciNet  Google Scholar 

  6. Fujishima, Y., Ishige, K.: Blowing up solutions for nonlinear parabolic systems with unequal elliptic operators. J. Dyn. Differ. Equ. 32, 1219–1231 (2020)

    Article  MathSciNet  Google Scholar 

  7. Fujishima, Y., Ishige, K.: Initial traces and solvability of Cauchy problem to a semilinear parabolic system. J. Math. Soc. Jpn. 73, 1187–1219 (2021)

    Article  MathSciNet  Google Scholar 

  8. Fujita, H.: On the blowing up of solutions of the Cauchy problem for \(u_{t}=\bigtriangleup u+ u^{1+\alpha}\). J. Fac. Sci., Univ. Tokyo, Sect. 1A, Math. 13, 109–124 (1966)

    Google Scholar 

  9. Giga, Y.: Surface Evolution Equations, A Level Set Approach. Monographs in Mathematics, vol. 99. Birkhäuser, Basel (2006)

    MATH  Google Scholar 

  10. Huang, Y., V’azquez, J.L.: Large-time geometrical properties of solutions of the Barenblatt equation of elasto-plastic filtration. J. Differ. Equ. 252, 4229–4242 (2012)

    Article  MathSciNet  Google Scholar 

  11. Imbert, C., Silvestre, L.: An Introduction to Fully Nonlinear Parabolic Equations, An Introduction to the Kähler-Ricci Flow. Lecture Notes in Math., vol. 2086, pp. 7–88. Springer, Cham (2013)

    MATH  Google Scholar 

  12. Ishii, H., Koike, S.: Viscosity solutions for monotone systems of second-order elliptic PDEs. Commun. Partial Differ. Equ. 16, 1095–1128 (1991)

    Article  MathSciNet  Google Scholar 

  13. Kamin, S., Peletier, L.A., Vázquez, J.L.: On the Barenblatt equation of elastoplastic filtration. Indiana Univ. Math. J. 40, 1333–1362 (1991)

    Article  MathSciNet  Google Scholar 

  14. Koike, S.: A beginner’s guide to the theory of viscosity solutions. In: MSJ Memoir, vol. 13 (2004). Math. Soc. Japan

    Google Scholar 

  15. Krylov, N.V.: Sobolev and Viscosity Solutions for Fully Nonlinear Elliptic and Parabolic Equations. Mathematical Surveys and Monographs, vol. 233. Am. Math. Soc., Providence (2018)

    Book  Google Scholar 

  16. Lieberman, G.M.: Second Order Parabolic Differential Equations. World Scientific Publishing Co., Inc., River Edge (1996)

    Book  Google Scholar 

  17. Meier, P.: On the critical exponent for reaction-diffusion equations. Arch. Ration. Mech. Anal. 109, 63–71 (1990)

    Article  MathSciNet  Google Scholar 

  18. Meneses, R., Quaas, A.: Fujita type exponent for fully nonlinear parabolic equations and existence results. J. Math. Anal. Appl. 376, 514–527 (2011)

    Article  MathSciNet  Google Scholar 

  19. Meneses, R., Quaas, A.: Existence and non-existence of global solutions for uniformly parabolic equations. J. Evol. Equ. 12, 943–955 (2012)

    Article  MathSciNet  Google Scholar 

  20. Uda, Y.: The critical exponent for a weakly coupled system of the generalized Fujita type reaction-diffusion equations. Z. Angew. Math. Phys. 46, 366–383 (1995)

    Article  MathSciNet  Google Scholar 

  21. Quittner, P., Souplet, Ph.: Superlinear Parabolic Problems, Blow-up, Global Existence and Steady States, 2nd edn. Springer, Cham (2019)

    Book  Google Scholar 

  22. Wang, L.: On the regularity theory of fully nonlinear parabolic equations: II. Commun. Pure Appl. Math. 45, 141–178 (1992)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgement

The authors thank the referee for his/her careful reading and valuable comments.

Funding

T. Kosugi is partially supported by Grant-in-Aid for Early-Career Scientists JSPS KAKENHI Grant Number 18K13436 and Tottori University of Environmental Studies Grant-in-Aid for Special Research. R. Sato is partially supported by Grant-in-Aid for Early-Career Scientists JSPS KAKENHI Grant Number 18K13435 and funding from Fukuoka University (Grant No. 205004).

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to this work.

Corresponding author

Correspondence to Takahiro Kosugi.

Ethics declarations

Competing Interests

Not applicable.

Additional information

Publisher’s Note

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

Appendix A: Perron’s Method

Appendix A: Perron’s Method

In this appendix, we state Perron’s method and give its proof. Let \(T>0\) and let \(S\) be a nonempty subset of viscosity subsolutions of (1.1) in \(\boldsymbol{R}^{N}\times (0,T)\).

Lemma A.1

Let \(T>0\). Assume that \(S\neq \emptyset \). For \(i=1,2\), set

$$ u_{i} (x,t):= \sup \{v_{i}(x,t) | \ v=(v_{1},v_{2}) \in S\}, \quad x \in \boldsymbol{R}^{N}, t\in (0,T). $$

If

$$\sup _{K} |u_{i}| < \infty ,\quad i=1,2 $$

for any compact subset \(K\subset \boldsymbol{R}^{N} \times (0,T)\), then \(u=(u_{1},u_{2})\) is a viscosity subsolution of (1.1).

Proof

For \(i=1,2\) and \(\varphi \in C^{2,1}(\boldsymbol{R}^{N}\times (0,T))\), we assume that \(u_{i}^{*}-\varphi \) attains its strict maximum at \((x_{0},t_{0})\in \boldsymbol{R}^{N} \times (0,T)\). Choose \(r>0\) so that \([t_{0}-r,t_{0}+r]\subset (0,T)\). Then for any \(\varepsilon >0\), we have

$$ u_{i}^{*}(x_{0},t_{0}) = \lim _{\varepsilon \to 0} \sup _{ \substack{x\in B(x_{0},\varepsilon ) \\ |t-t_{0}|< \varepsilon }} u_{i}(x,t) \leq \sup _{ \substack{x\in B(x_{0},\varepsilon ) \\ |t-t_{0}|< \varepsilon }} u_{i}(x,t), $$

where \(u_{i}^{*}\) us defined in (2.1). For all \(\tau >0\), there exist sequences \(x_{\tau ,\varepsilon }\in B(x_{0},\varepsilon )\) and \(t_{\tau ,\varepsilon }\in (t_{0}-r, t_{0}+r)\) such that

$$ \sup _{ \substack{x\in B(x_{0},\varepsilon ) \\ |t-t_{0}|< \varepsilon }} u_{i}(x,t) \leq u_{i}(x_{\tau ,\varepsilon },t_{\tau ,\varepsilon }) +\tau . $$

For any \(\tau >0\) there exists \(\delta _{\tau }>0\) such that for all \(x\in \boldsymbol{R}^{N}\), \(t\in (0,T)\), if \(|x-x_{0}| + |t-t_{0}|<\delta _{\tau}\), then \(|\varphi (x,t) - \varphi (x_{0},t_{0})|<\tau \). For each \(k=1,2,\dots \) set

$$ \varepsilon =\min \left \{ \frac{\delta _{1/k}}{2}, \frac{1}{k},r \right \}. $$

There exist \(x_{k} \in B_{r}(x_{0})\), \(t_{k}\in (t_{0}-r,t+r)\) such that

$$ \begin{aligned} &x_{k} \to x_{0}, \quad t_{k}\to t_{0} \quad \mathrm{as} \ k\to \infty , \\ &u_{i}^{*}(x_{0},t_{0}) < u_{i}(x_{k},t_{k}) + \frac{1}{k}, \quad | \varphi (x_{k},t_{k})- \varphi (x_{0},t_{0})| < \frac{1}{k}. \end{aligned} $$

Moreover, by the definition of \(u_{i}\), there exists \((u_{1}^{k},u_{2}^{k})\in S\) such that

$$ u_{i}(x_{k},t_{k}) < u_{i}^{k}(x_{k},t_{k}) +\frac{1}{k}. $$

Choose \((y_{k},s_{k})\in \overline{B}_{r}(x_{0})\times [t_{0}-r,t_{0}+r]\) so that \((u_{i}^{k})^{*} -\varphi \) attains its maximum at \((y_{k},s_{k})\). Taking a subsequence (still denoted by the same symbol), we see that as \(k\to \infty \), \(y_{k} \to \hat{y}\), \(t_{k}\to \hat{t}\) for some \(\hat{y}\in \overline{B}_{r}(x_{0})\), \(t_{k}\to \hat{t}\in [t_{0}-r,t_{0}+r]\). We have

$$ \begin{aligned} (u_{i}^{*}-\varphi )(x_{0},t_{0}) & < (u_{i}^{k}-\varphi )(x_{k},t_{k}) +\frac{3}{k} \\ &\leq ((u_{i}^{k})^{*}-\varphi )(x_{k},t_{k}) +\frac{3}{k} \\ &\leq ((u_{i}^{k})^{*}-\varphi )(y_{k},s_{k}) +\frac{3}{k} \\ &\leq (u_{i}^{*}-\varphi )(y_{k},s_{k}) +\frac{3}{k}. \end{aligned} $$
(A.1)

Since \(u_{i}^{*}\) is upper semicontinuous, we see that

$$ (u_{i}^{*} -\varphi )(x_{0},t_{0}) \leq (u_{i}^{*} - \varphi )( \hat{y},\hat{t}). $$

On the other hand, \(u_{i}^{*} -\varphi \) has a strict maximum at \((x_{0},t_{0})\), hence \(x_{0}=\hat{y}\) and \(t_{0} =\hat{t}\). Therefore, \(y_{k}\in B_{r}(x_{0})\), \(t_{k}\in (t_{0}-r,t_{0}+r)\) for sufficiently large \(k\). In addition, using (A.1) again, we have

$$ \lim _{k\to \infty}(u_{i}^{k})^{*}(y_{k},s_{k}) = u_{i}^{*}(x_{0},t_{0}) $$

and

$$ \begin{aligned} \limsup _{k\to \infty} (u_{i}^{k})^{*}(y_{k},s_{k}) \leq \limsup _{k \to \infty}u_{j}^{*}(y_{k},s_{k}) \leq u_{j}^{*}(x_{0},t_{0}), \end{aligned} $$

where \(j\neq i\). Consequently, we obtain

$$ \partial _{t} \varphi (y_{k},s_{k}) + F_{i}(y_{k},D^{2}\varphi (y_{k},s_{k})) \leq |u_{j}^{*}(y_{k},s_{k})|^{p_{i} -1}u_{j}^{*}(y_{k},s_{k}), $$

hence

$$ \partial _{t} \varphi (x_{0},t_{0}) + F_{i}(x_{0},D^{2}\varphi (x_{0},t_{0})) \leq |u_{j}^{*}(x_{0},t_{0})|^{p_{i} -1}u_{j}^{*}(x_{0},t_{0}), $$

which completes the proof. □

Proposition A.1

Assume that \(\xi =(\xi _{1},\xi _{2})\in (L^{\infty}_{\mathrm{loc}}(\boldsymbol{R}^{N} \times (0,T)))^{2}\) is a viscosity subsolution of (1.1) and \(\eta =(\eta _{1},\eta _{2})\in (L^{\infty}_{\mathrm{loc}}(\boldsymbol{R}^{N} \times (0,T)))^{2}\) is a viscosity supersolution of (1.1) for some \(T>0\). If \(\xi \) and \(\eta \) satisfy

$$\xi _{1} \leq \eta _{1}, \quad \xi _{2} \leq \eta _{2} \quad \mathrm{in} \ \boldsymbol{R}^{N} \times (0,T), $$

then

$$ u_{i}(x,t) := \sup \{v_{i}(x,t) \mid v=(v_{1},v_{2})\in S, \ \xi \leq v \leq \eta \} $$

is a viscosity solution of (1.1) in \(\boldsymbol{R}^{N}\times (0,T)\). Here \(\xi \leq v \leq \eta \) means that \(\xi _{i} \leq v_{i} \leq \eta _{i}\) in \(\boldsymbol{R}^{N}\times (0,T)\) for \(i=1,2\).

Proof

By Lemma A.1, \(u=(u_{1},u_{2})\) is a viscosity subsolution to (1.1). Suppose, contrary to our claim, that there exist \(\varphi \in C^{2,1}\) and \((x_{0},t_{0})\in \boldsymbol{R}^{N}\times (0,T)\), \(u_{i*} -\varphi \) attains a strict minimum at \((x_{0},t_{0})\), \((u_{i*}-\varphi )(x_{0},t_{0}) =0\) for some \(i=1,2\) and exists \(\theta >0\) such that

$$ \partial _{t} \varphi + F_{i}(x_{0},D^{2}\varphi ) - |u_{j*}|^{p_{i} -1}u_{j*} < -\theta $$
(A.2)

at \((x_{0},t_{0})\), where \(i\neq j\) and \(p_{1} =p\), \(p_{2} =q\).

We firstly show that \(\varphi (x_{0},t_{0}) <\eta _{i*}(x_{0},t_{0})\). In fact, we see that \(\varphi \leq u_{i*} \leq \eta _{i*}\) and \(u_{j*} \leq \eta _{j*}\) and \(\eta _{i*} - \varphi \) attains a minimum at \((x_{0},t_{0})\) if \(\varphi (x_{0},t_{0}) = \eta _{i*}(x_{0},t_{0})\), thus by the definition of the viscosity supersolution, we obtain

$$ \partial _{t} \varphi (x_{0},t_{0}) + F(x_{0},D^{2}\varphi (x_{0},t_{0})) \geq |\eta _{j*}|^{p_{j} - 1}\eta _{j*} (x_{0},t_{0}) \geq |u_{j*}|^{p_{j} -1}u_{j*}(x_{0},t_{0}). $$

This contradicts the assumption.

For any \(\rho >0\), there exists \(\varepsilon _{\rho}>0\) such that

$$ \begin{aligned} u_{j*}(x_{0},t_{0}) -\rho &= \sup _{\varepsilon >0}\inf _{ \substack{|x-x_{0}|< \varepsilon \\|t-t_{0}|< \varepsilon }} u_{j}(x,t) - \rho \\ &< \inf _{ \substack{|x-x_{0}|< \varepsilon _{\rho }\\|t-t_{0}|< \varepsilon _{\rho}}} u_{j}(x,t) \\ &\leq u_{j}(x,t) \end{aligned} $$

for \(|x-x_{0}|<\varepsilon _{\rho}\) and \(|t-t_{0}|<\varepsilon _{\rho}\). By using the mean value theorem, there exists \(\hat{\theta}\) such that

$$ \begin{aligned} |u_{j*}|^{p_{i} -1}u_{j*} &\geq |u_{j*}(x_{0},t_{0}) -\rho |^{p_{i} -1}(u_{j*}(x_{0},t_{0})- \rho ) \\ &= |u_{j*}(x_{0},t_{0}) |^{p_{i} -1}u_{j*}(x_{0},t_{0}) \\ &- \rho p_{i} | \hat{\theta} u_{j*}(x_{0},t_{0})(x_{0},t_{0}) + (1- \hat{\theta})(u_{j*}(x_{0},t_{0})-\rho ) |^{p_{i} -1}. \end{aligned} $$

We can find \(\rho >0\) and \(s_{0}>0\) so small that

$$ \rho p_{i} | \hat{\theta} u_{j*}(x_{0},t_{0})(x_{0},t_{0}) + (1- \hat{\theta})(u_{j*}(x_{0},t_{0})-\rho ) |^{p_{i} -1} < \frac{\theta}{4} $$

and

$$ \vert \partial _{t}\varphi (x_{0},t_{0}) + F_{i}(x_{0},D^{2}\varphi (x_{0},t_{0})) - \partial _{t}\varphi (x,t) - F_{i}(x,D^{2}\varphi (x,t)) \vert < \frac{\theta}{4} $$

for \(|x-x_{0}|< s_{0}\) and \(|t-t_{0}|< s_{0}\). This together with (A.2) and the continuity of \(\partial _{t}\varphi \) and \(F_{i}(\cdot ,D^{2}\varphi )\) implies that

$$ \begin{aligned} -\theta &> \partial _{t} \varphi (x,t) +F_{i}(x,D^{2}\varphi (x,t)) - \frac{\theta}{4} -|u_{j*}|^{p_{i}-1}u_{j*}(x,t) -\frac{\theta}{4} \end{aligned} $$

for \(|x-x_{0}| < s_{0}\) and \(|t-t_{0}|< s_{0}\). Therefore,

$$ \partial _{t} \varphi (x,t) +F_{i}(x,D^{2}\varphi (x,t)) -|u_{j*}|^{p_{i}-1}u_{j*}(x,t) < -\frac{\theta}{2}. $$
(A.3)

It is already shown that \(u_{i*}(x_{0},t_{0})=\varphi (x_{0},t_{0}) <\eta _{i*}(x_{0},t_{0})\). Set

$$ 3\hat{\tau} :=\eta _{i*}(x_{0},t_{0}) -u_{i*}(x_{0},t_{0})>0. $$

Since \(\eta _{i*}\) is lower semicontinuous and \(\varphi \) is continuous, we can find \(s_{1}\in (0,s_{0})\) such that for all \(|x-x_{0}| < s_{1}\) and \(t\in (t_{0} -s_{1},t_{0}+s_{1})\),

$$ \eta _{i*}(x,t) -\varphi (x,t) > \eta _{i*}(x_{0},t_{0}) -\varphi (x_{0},t_{0}) - \hat{\tau} = 2 \hat{\tau}. $$

Therefore \(\varphi (x,t) +2\hat{\tau} < \eta _{i*}(x,t)\) in \(D\), where

$$ D:= B_{s_{1}}(x_{0}) \times (t_{0}-s_{1},t_{0}+s_{1}). $$

On the other hand, since \(u_{i*}-\varphi \) attains a strict minimum at \((x_{0},t_{0})\) and \((u_{i*}-\varphi )(x_{0},t_{0})=0\), there exist \(\varepsilon \in (0,s_{1} /2)\) and \(\tau _{0} \in (0,\hat{\tau})\) such that

$$ (u_{i*}-\varphi )(x,t) \geq \min _{(x,t)\in A}\{ (u_{i*}-\varphi )(x,t) \} > \tau _{0}. $$

Here we have set

$$ A := \left ( \overline{B_{s_{1} /2 +\varepsilon }(x_{0})}\setminus B_{s_{1} /2 -\varepsilon }(x_{0}) \right ) \times \left \{ t \mid \frac{s_{1}}{2}-\varepsilon \leq |t-t_{0}| \leq \frac{s_{1}}{2} + \varepsilon \right \}. $$

We now define \((w_{1},w_{2})\) by

$$ \begin{aligned} w_{i}(x,t) &:= \textstyle\begin{cases} \max \{u_{i}(x,t),\varphi (x,t) + \tau _{0}\} \quad &\mathrm{in} \ D/2, \\ u_{i}(x,t) \quad \mathrm{in} \ (\boldsymbol{R}^{N}\times (0,T)) \setminus (D/2), \end{cases}\displaystyle \\ w_{j} (x,t) &:= u_{j}(x,t) \quad \mathrm{in} \ \boldsymbol{R}^{N}\times (0,T), \end{aligned} $$

where

$$ D/2 := B_{s_{1} /2}(x_{0}) \times \left (t_{0}- \frac{s_{1}}{2},t_{0}+ \frac{s_{1}}{2} \right ). $$

In what follows, we shall show that \((w_{1},w_{2})\) is a viscosity subsolution to (1.1) in \(\boldsymbol{R}^{N}\times (0,T)\) satisfying \(\xi _{k} \leq w_{k} \leq \eta _{k}\) for \(k=1,2\). It follows from the definition of \(w_{i}\) that we have \(\xi _{i} \leq u_{i} \leq w_{i}\) in \(\boldsymbol{R}^{N}\times (0,T)\). Since \(\varphi (x,t) + \tau _{0} \leq \eta _{i*}\) in \(D\), we see that \(\varphi + \tau _{0} \leq \eta _{i}\) in \(D\), hence \(w_{k} \leq \eta _{k}\) in \(\boldsymbol{R}^{N}\times (0,T)\) for \(k=1,2\). Consequently, for \(k=1,2\), we obtain

$$ \xi _{k} \leq w_{k} \leq \eta _{k} \quad \mathrm{in} \ \boldsymbol{R}^{N} \times (0,T). $$

We can find \(n\in \{1,2,\dots \}\) sufficiently large so that

$$ \frac{1}{n} < \frac{s_{1}}{2} -\varepsilon \quad \mathrm{and} \quad \frac{1}{n} < \frac{\tau _{0}}{2} $$

and there exist \(x_{n} \in B_{1/n}(x_{0})\) and \(t_{n}\in \boldsymbol{R}\) with \(|t_{0}-t_{n}|<1/n\) such that

$$ u_{i} (x_{n},t_{n}) < u_{i*}(x_{0},t_{0}) + \frac{1}{n}. $$

Moreover, it follows that

$$ u_{i*}(x_{0},t_{0}) + \frac{1}{n} < u_{i*}(x_{0},t_{0}) + \frac{\tau _{0}}{2} = \varphi (x_{0},t_{0}) + \frac{\tau _{0}}{2} < \varphi (x_{0},t_{0}) + \tau _{0}. $$

Note that \((x_{n},t_{n}) \in D/2\).

We next prove that \((w_{1},w_{2})\) is a viscosity subsolution to (1.1) in \(\boldsymbol{R}^{n}\times (0,T)\). Let us take \(\boldsymbol{R}^{N}\times (0,T)\) and \(\psi \in C^{2,1}(\boldsymbol{R}^{N}\times (0,T))\) arbitrarily.

We firstly assume that \(w_{i}^{*} - \psi \) attains a local maximum at \((\hat{x},\hat{t})\). Consider the first case \(w_{i}^{*} (\hat{x},\hat{t}) = u_{i}^{*} (\hat{x},\hat{t})\). Then

$$ \begin{aligned} u_{i}^{*} (\hat{x},\hat{t}) -\psi (\hat{x},\hat{t}) &= w_{i}^{*}( \hat{x},\hat{t}) - \psi (\hat{x},\hat{t}) \\ &\geq w_{i}^{*} (x,t) -\psi (x,t) \\ &\geq u_{i}^{*} (x,t) - \psi (x,t) \end{aligned} $$

in \(\boldsymbol{R}^{N}\times (0,T)\). Thus, \(u_{i}^{*} -\psi \) attains its maximum at \((\hat{x},\hat{t})\). Moreover, since \((u_{1},u_{2})\) is a subsolution to (1.1) in \(\boldsymbol{R}^{N}\times (0,T)\) and \(u_{j} \equiv w_{j}\), we have

$$\partial _{t} \psi + F_{i}(\cdot ,D^{2}\psi ) \leq |u_{j}^{*}|^{p_{i} -1}u_{j}^{*} =|w_{j}^{*}|^{p_{i} -1}w_{j}^{*} $$

at \((\hat{x},\hat{t})\).

We now consider the second case \(w_{i}^{*}(\hat{x},\hat{t}) =(\varphi +\tau _{0})^{*}(\hat{x},\hat{t}) =\varphi (\hat{x},\hat{t})+\tau _{0}\). Note that \((\hat{x},\hat{t})\in D/2\). The same argument above implies that \(\varphi +\tau _{0}-\psi \) attains its maximum at \((\hat{x},\hat{t})\). Thus, we see that

$$ \partial _{t} \varphi (\hat{x},\hat{t}) = \partial _{t} \psi (\hat{x}, \hat{t}), \quad D\varphi (\hat{x},\hat{t}) = D\psi (\hat{x},\hat{t}), \quad D^{2} \varphi (\hat{x},\hat{t}) \leq D^{2} \psi (\hat{x}, \hat{t}). $$

It follows from (1.8) and (A.3) that

$$ \begin{aligned} \partial _{t} \psi + F_{i}(\cdot ,D^{2}\psi ) &\leq \partial _{t} \varphi + F_{i}(\cdot ,D^{2}\varphi ) \\ &\leq |{u_{j}}_{*}|^{p_{i} -1} {u_{j}}_{*} \\ &\leq |{u_{j}}^{*}|^{p_{i} -1} {u_{j}}^{*} \\ &= |{w_{j}}^{*}|^{p_{i} -1} {w_{j}}^{*} \end{aligned} $$

at \((\hat{x},\hat{t})\).

We secondly assume that \(w_{j}^{*} - \psi \) attains a local maximum at \((\hat{x},\hat{t})\). Since \(w_{j} =u_{j}\) in \(\boldsymbol{R}^{N}\times (0,T)\), \(u_{j}^{*}-\psi \) also attains its maximum at \((\hat{x},\hat{t})\). Therefore, we obtain

$$ \begin{aligned} \partial _{t} \psi + F_{j}(\cdot ,D^{2}\psi ) \leq |u_{i}^{*}|^{p_{j} -1}u_{i}^{*} \leq |w_{i}^{*}|^{p_{j} -1}w_{i}^{*} \end{aligned} $$

at \((\hat{x},\hat{t})\).

Consequently, \((w_{1},w_{2})\) is a viscosity subsolution to (1.1) in \(\boldsymbol{R}^{N}\times (0,T)\). This contradicts the definition of \((u_{1},u_{2})\). □

Rights and permissions

Springer Nature or its licensor 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

Kosugi, T., Sato, R. Existence of Global-in-Time Solutions to a System of Fully Nonlinear Parabolic Equations. Acta Appl Math 181, 14 (2022). https://doi.org/10.1007/s10440-022-00533-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10440-022-00533-7

Keywords

Mathematics Subject Classification

Navigation