Skip to main content
Log in

Invariant measures for place-dependent idempotent iterated function systems

  • Published:
Journal of Fixed Point Theory and Applications Aims and scope Submit manuscript

Abstract

We study the set of invariant idempotent probabilities for place-dependent idempotent iterated function systems defined in compact metric spaces. Using well-known ideas from dynamical systems, such as the Mañé potential and the Aubry set, we provide a complete characterization of the densities of such idempotent probabilities. As an application, we provide an alternative formula for the attractor of a class of fuzzy iterated function systems.

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

Notes

  1. For any sets A and B and any function \(f:A\times B \rightarrow \mathbb {R}_{\max }\) bounded from above, by definition of supremum, we have   \(\displaystyle {\bigoplus _{a\in A}\left[ \bigoplus _{b\in B}f(a,b)\right] = \bigoplus _{(a,b)\in A\times B}f(a,b) = \bigoplus _{b\in B}\left[ \bigoplus _{a\in A}f(a,b)\right] .}\)

References

  1. Akian, M.: Densities of idempotent measures and large deviations. Trans. Am. Math. Soc. 351(11), 4515–4543 (1999)

    Article  MathSciNet  Google Scholar 

  2. Akian, M., Quadrat, J.-P., Viot, M.: Duality between probability and optimization. In: Idempotency (Bristol, 1994), Publ. Newton Inst., vol. 11. Cambridge University Press, Cambridge, pp. 331–353 (1998)

  3. Baraviera, A., Leplaideur, R., Lopes, A.: Ergodic optimization, zero temperature limits and the max-plus algebra. Publicações Matemáticas do IMPA. [IMPA Mathematical Publications]. Instituto Nacional de Matemática Pura e Aplicada (IMPA), Rio de Janeiro. 29\(^{\rm o}\) Colóquio Brasileiro de Matemática. [29th Brazilian Mathematics Colloquium] (2013)

  4. Baraviera, A., Lopes, A.O., Thieullen, P.: A large deviation principle for the equilibrium states of Hölder potentials: the zero temperature case. Stoch. Dyn. 6(1), 77–96 (2006)

    Article  MathSciNet  Google Scholar 

  5. Bazylevych, L., Repovš, D., Zarichnyi, M.: Spaces of idempotent measures of compact metric spaces. Topol. Appl. 157(1), 136–144 (2010)

    Article  MathSciNet  Google Scholar 

  6. Cabrelli, C.A., Forte, B., Molter, U.M., Vrscay, E.R.: Iterated fuzzy set systems: a new approach to the inverse problem for fractals and other sets. J. Math. Anal. Appl. 171(1), 79–100 (1992)

    Article  MathSciNet  Google Scholar 

  7. Contreras, G., Lopes, A.O., Thieullen, P.: Lyapunov minimizing measures for expanding maps of the circle. Ergod. Theory Dyn. Syst. 21(5), 1379–1409 (2001)

    Article  MathSciNet  Google Scholar 

  8. Conze, J.-P., Guivarc’h, Y., Propriété.: de droite fixe et fonctions propres des opérateurs de convolution. In: Séminaire de Probabilités, I (Univ. Rennes, Rennes, 1976). Université de Rennes, Département de Mathématiques et Informatique, Rennes, Exp. No. 4, 22 (1976)

  9. da Cunha, R.D., Oliveira, E.R., Strobin, F.: Existence of invariant idempotent measures by contractivity of idempotent Markov operators. J. Fixed Point Theory Appl. 25(1), Paper No. 8, 11 (2023)

  10. da Cunha, R.D., Oliveira, E.R., Strobin, F.: Fuzzy-set approach to invariant idempotent measures. Fuzzy Sets Syst. 457, 46–65 (2023)

    Article  MathSciNet  Google Scholar 

  11. Del Moral, P., Doisy, M.: Maslov idempotent probability calculus. I. Teor. Veroyatnost. i Primenen. 43(4), 735–751 (1998)

    Article  MathSciNet  Google Scholar 

  12. Diamond, P., Kloeden, P.: Metric spaces of fuzzy sets. World Scientific Publishing Co., Inc., River Edge (1994). Theory and applications

  13. Fan, A.H., Lau, K.-S.: Iterated function system and Ruelle operator. J. Math. Anal. Appl. 231(2), 319–344 (1999)

    Article  MathSciNet  Google Scholar 

  14. Garibaldi, E., Lopes, A.O.: On the Aubry–Mather theory for symbolic dynamics. Ergod. Theory Dyn. Syst. 28(3), 791–815 (2008)

    Article  MathSciNet  Google Scholar 

  15. Iturriaga, R., Lopes, A.O., Mengue, J.K.: Selection of calibrated subaction when temperature goes to zero in the discounted problem. Discrete Contin. Dyn. Syst. 38(10), 4997–5010 (2018)

    Article  MathSciNet  Google Scholar 

  16. Kolokol’tsov, V.N., Maslov, V.P.: Idempotent analysis as a technique of the theory of control and optimal synthesis. I. Funktsional. Anal. i Prilozhen. 23(1), 1–14 (1989)

    Article  MathSciNet  Google Scholar 

  17. Kolokol’tsov, V.N., Maslov, V.P.: Idempotent analysis as a technique of the theory of control and optimal synthesis. II. Funktsional. Anal. i Prilozhen 23(4), 53–62, 92 (1989)

    MathSciNet  Google Scholar 

  18. Kolokoltsov, V.N., Maslov, V.P.: Idempotent analysis and its applications, Mathematics and its Applications, vol. 401. Kluwer Academic Publishers Group, Dordrecht (1997). Translation of Idempotent analysis and its application in optimal control (Russian), “Nauka” Moscow, [MR1375021 (97d:49031)], Translated by V. E. Nazaikinskii, With an appendix by Pierre Del Moral (1994)

  19. Leplaideur, R., Mengue, J.: On the selection of subaction and measure for perturbed potentials, preprint (2024). arXiv:2404.02182v1

  20. Lewellen, G.B.: Self-similarity. Rocky Mt. J. Math. 23(3), 1023–1040 (1993)

    Article  MathSciNet  Google Scholar 

  21. Litvinov, G.L., Maslov, V.P.: Idempotent mathematics: the correspondence principle and its computer realizations. Uspekhi Mat. Nauk 51 6(312), 209–210 (1996)

    MathSciNet  Google Scholar 

  22. Mañé, R.: Lagrangian flows: the dynamics of globally minimizing orbits. Bol. Soc. Brasil. Mat. (N.S.) 28(2), 141–153 (1997)

    Article  MathSciNet  Google Scholar 

  23. Mazurenko, N., Zarichnyi, M.: Invariant idempotent measures. Carpathian Math. Publ. 10(1), 172–178 (2018)

  24. Mengue, J.: Large deviations for equilibrium measures and selection of subaction. Bull. Braz. Math. Soc. 49(1), 17–42 (2018)

    Article  MathSciNet  Google Scholar 

  25. Mengue, J., Oliveira, E.: Duality results for iterated function systems with a general family of branches. Stoch. Dyn. 17(03), 1750021 (2017)

    Article  MathSciNet  Google Scholar 

  26. Mihail, A.: The shift space of a recurrent iterated function system. Rev. Roumaine Math. Pures Appl. 53(4), 339–355 (2008)

    MathSciNet  Google Scholar 

  27. Mihail, A., Miculescu, R.: Applications of fixed point theorems in the theory of generalized IFS. Fixed Point Theory Appl. 11, 312876 (2008)

    Article  MathSciNet  Google Scholar 

  28. Oliveira, E.R., Strobin, F.: Fuzzy attractors appearing from GIFZS. Fuzzy Sets Syst. 331, 131–156 (2018)

    Article  MathSciNet  Google Scholar 

  29. Phelps, R.R.: Lectures on Choquet’s theorem, 2nd edn. Lecture Notes in Mathematics, vol. 1757. Springer, Berlin (2001)

  30. Strobin, F., Swaczyna, Ja.: On a certain generalisation of the iterated function system. Bull. Aust. Math. Soc. 87(1), 37–54 (2013)

    Article  MathSciNet  Google Scholar 

  31. Zadeh, L.A.: Fuzzy sets. Inf Control 8, 338–353 (1965)

    Article  Google Scholar 

  32. Zaitov, A.A.: On a metric on the space of idempotent probability measures. Appl. Gen. Topol. 21(1), 35–51 (2020)

    Article  MathSciNet  Google Scholar 

  33. Zarichnyĭ, M.M.: Spaces and mappings of idempotent measures. Izv. Ross. Akad. Nauk Ser. Mat. 74(3), 45–64 (2010)

    MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Elismar R. Oliveira.

Additional information

Publisher's Note

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

Appendix A: Fundamentals of idempotent analysis

Appendix A: Fundamentals of idempotent analysis

The following exposition is based on the ideas of Kolokol’tsov and Maslov [16], who considered a different setting for applications of idempotent analysis. See also Refs. [5, 9, 10, 17, 18, 21, 23, 32, 33], among many others for additional references.

1.1 Proof of Theorem 1.2

The proof of Theorem 1.2 will be divided in several parts as exposed below.

Proposition 6.1

Any \(m\in C^*(X,\mathbb {R})\) is order preserving, that is, if \(f \le g\) then \(m( f) \le m( g)\). Moreover

$$\begin{aligned} \min _{X} f \le m(f) - m(0) \le \max _{X} f. \end{aligned}$$

Proof

If \(f \le g\) then \(f \oplus g = g\). As m is max-plus additive, we obtain \(m(g) = m(f \oplus g)=m( f) \oplus m( g)\), thus \(m( f) \le m( g)\).

As we have a function \(m: C(X,\mathbb {R}) \rightarrow \mathbb {R}\), then we get \(m(0)\ne -\infty \). As for any \(f \in C(X, \mathbb {R})\), \(\min _{X} f \le f(x) \le \max _{X} f\), we obtain

$$\begin{aligned}{} & {} m(\min _{X} f ) \le m(f) \le m(\max _{X} f) \\{} & {} m((\min _{X} f)\odot 0) \le m(f) \le m((\max _{X} f)\odot 0) \\{} & {} (\min _{X} f) \odot m(0) \le m(f) \le (\max _{X} f) \odot m( 0). \end{aligned}$$

\(\square \)

The semimodule \(\mathcal {V}:=(C(X,\mathbb {R}), \oplus , \odot )\) can be topologized with the usual structure given by the metric \(d_{\infty }(f,g)=\bigoplus _{x\in X} |f(x)- g(x)|\). Introducing a metric \(\rho : \mathbb {R}_{\max } \times \mathbb {R}_{\max } \rightarrow [0, +\infty )\) given by \(\rho (a,b):=|\exp (a)-\exp (b)|\), we obtain a topological semiring \((\mathbb {R}_{\max }, \rho )\) and for the set \(\mathcal {F}(X,\mathbb {R}_{\max }):=\{ f: X \rightarrow \mathbb {R}_{\max }\}\), we can consider the topology given by

$$\begin{aligned} d_{\rho }(f,g)=\bigoplus _{x\in X} \rho (f(x), g(x))=\bigoplus _{x\in X} |\exp (f(x))- \exp (g(x))|, \end{aligned}$$

for any \(f, g \in \mathcal {F}(X,\mathbb {R}_{\max })\).

Proposition 6.2

Any \(m\in C^*(X,\mathbb {R})\) is nonexpansive with respect to the usual sup-norm \(d_{\infty }\) in \(C(X, \mathbb {R})\) and the absolute value \(|\cdot |\) in \(\mathbb {R}\). In particular, it is continuous.

Proof

Indeed,

$$\begin{aligned}&\displaystyle -d_{\infty }(f, g) + g(x) \le f(x) \le d_{\infty }(f, g) + g(x), \forall x \in X\\&\displaystyle -d_{\infty }(f, g) + m(g) \le m(f) \le d_{\infty }(f, g) + m(g)\\&\displaystyle |m(f) - m(g)| \le d_{\infty }(f, g). \end{aligned}$$

\(\square \)

Definition 6.3

We say that a sequence of functions \(f_n \in C(X,\mathbb {R})\) converges pointwise to a function \(f \in \mathcal {F}(X,\mathbb {R}_{\max })\) if \(\lim _{n\rightarrow \infty } \rho (f_n(x),f(x)) =0\), for any \(x\in X\). Equivalently:

  1. 1.

    \(\lim _{n\rightarrow +\infty }f_n(x) = f(x)\), for all \(x\in X\) such that \(f(x)\in \mathbb {R} \)

  2. 2.

    \(\lim _{n\rightarrow +\infty }f_n(x) = -\infty \), for all \(x\in X\) such that \(f(x)=-\infty . \)

Definition 6.4

Given \(a \in \mathbb {R}\) and \(x \in X\), a fixed point, we define the Dirac function

$$\begin{aligned} g^{a}_{x}(y):= \left\{ \begin{array}{ll} a, &{}\quad y=x \\ -\infty , &{}\quad y\ne x \end{array} \right. \end{aligned}$$

and \(\Delta (X, \mathbb {R})\) the set of these functions.

Lemma 6.5

There exists a monotone nonincreasing sequence of continuous functions \(f_n\in C(X,\mathbb {R})\) which converges pointwise to \(g^{a}_{x}\).

Proof

A consequence of the fact that X is compact metric space is that we can easily build a sequence in \(C(X, \mathbb {R})\) of nonincreasing functions converging pointwise to \(g^{a}_{x}\). Indeed, for each \(n \in \mathbb {N}\), consider the function \(f_{n}\) defined by

$$\begin{aligned} f_{n}(y):= \left\{ \begin{array}{ll} a, &{}\quad d(y,x)<1/n \\ -n, &{}\quad d(y,x)>2/n \\ a-(n^2 + a n)(d(y,x) - 1/n), &{}\quad 1/n \le d(y,x) \le 2/n \end{array} \right. . \end{aligned}$$
(16)

It is obviously continuous in y and monotone nonincreasing in n. Moreover, for any \(y \in X\), \(y\ne x\), we get \(d(y,x)>2/n\) for n large enough. In this case, we obtain \(f_n(y)\rightarrow -\infty \). This proves the pointwise convergence. \(\square \)

Definition 6.6

Given \(g^{a}_{x}\), we will call the sequence of functions \((f_{n})\) given in (16) as the standard continuous approximation of \(g^{a}_{x}\).

Lemma 6.7

Let \((f_n)\) be a nonincreasing sequence of continuous functions converging pointwise to a continuous function f in a compact set X. Then the convergence is uniform.

Proof

Suppose in contradiction there is an \(\varepsilon >0\) and a sequence of points \((x_n)\in X\) such that \(f_n(x_n)>f(x_n)+\varepsilon ,\; \forall \,n\in \mathbb {N}\). As X is compact, there exists a subsequence \((x_{n_i})\) and a point \(x_0\in X\) such that \(x_{n_i}\rightarrow x_0\). Let \(n_j\) be such that \(f_{n_j}(x_0)<f(x_0)+\varepsilon /2\) (it exists because we have convergence pointwise). As f and \(f_{n_j}\) are continuous in \(x_0\), there exists a \(\delta >0\) such that \(d(x,x_0)<\delta \Rightarrow f_{n_j}(x)<f(x)+\varepsilon \). As \((f_n)\) is nonincreasing, we also get

$$\begin{aligned} (d(x,x_0)<\delta ,\,n\ge n_j) \Rightarrow f_{n}(x)<f(x)+\varepsilon . \end{aligned}$$

This is a contradiction because, by definition of \(x_n\) and \(x_0\), for n large enough, we must have \(d(x_n,x_0)<\delta \) and also \(f_n(x_n)>f(x_n)+\varepsilon \). \(\square \)

Lemma 6.8

Let \(\bar{C}(X,\mathbb {R})\) be the set of functions \(g:X\rightarrow \mathbb {R}_{\max }\) such that there exists a nonincreasing sequence of continuous functions \((f_n)\) converging pointwise to g. This set is closed with respect to \((\oplus ,\odot )\) operations. Given \(m \in C^{*}(X, \mathbb {R})\) an idempotent measure, we can extend it to an idempotent measure on \(\bar{C}(X,\mathbb {R})\) by \(\tilde{m}(g) = \lim _{n\rightarrow +\infty }m(f_n)\), where \((f_n)\) is any nonincreasing sequence of continuous functions converging pointwise to g.

Furthermore, if C is a set which is closed with respect to \((\oplus ,\odot )\) operations, such that \(C(X,\mathbb {R})\subseteq C \subseteq \bar{C}(X,\mathbb {R})\) and \(\overline{m}\) is an idempotent measure on C which extends m, then \(\overline{m}\le \tilde{m}\).

Proof

If \(g,g'\in \bar{C}(X,\mathbb {R})\) and \(f_n\rightarrow g,\,f'_n\rightarrow g'\), we have \(f_n \oplus f'_n \rightarrow g\oplus g'\) and \(a\odot f_n \rightarrow a\odot g\) then, \(\bar{C}(X,\mathbb {R}) \) is closed with respect to \((\oplus ,\odot )\) operations.

Consider any sequence of monotonous nonincreasing continuous functions \((f_{n})\) converging pointwise to g. We claim that the value \(\tilde{m}(g):=\lim _{n \rightarrow \infty } m(f_{n}) \in \mathbb {R}_{\max }\) is well defined and furthermore it does not depend on \((f_n)\).

Indeed, as \(f_{n} \ge f_{n+1}\), we have \(m(f_{n}) \ge m(f_{n+1})\). Thus, there exists the limit \(\lim _{n \rightarrow \infty } m(f_{n}) \in \mathbb {R}_{\max }\). Let \((f_{n})\) and \((f'_{n})\) be sequences of monotonous nonincreasing continuous functions converging pointwise to g. For any fixed k, we define a new sequence \(\psi _{n}:=f_{k}\oplus f'_{n},\) so that \(\psi _{n}\) converges pointwise to \(f_{k} \in C(X,\mathbb {R})\). Applying Lemma 6.8, we get that \(\psi _n\) converges uniformly to \(f_k\). From Proposition 6.2, we obtain \(m(f_k) = \lim _{n \rightarrow \infty } m(\psi _{n})\). Then \(m(f_k)= \lim _{n \rightarrow \infty } m(f_k \oplus f'_{n}) \ge \lim _{n \rightarrow \infty } m(f'_{n}).\) Now we can take the limit on the left hand side obtaining \(\lim _{k\rightarrow \infty }m(f_k) \ge \lim _{n \rightarrow \infty } m(f'_{n}).\) Reversing the role of the sequences, we obtain that the limits are equal.

It is easy to see that \(\tilde{m}\) is actually an extension of m because, for \(h \in C(X, \mathbb {R})\), we can take the sequence \(f_{n}=h, \forall n\), which is continuous, monotone, and not increasing, so \(\tilde{m}(h)= \lim _{n \rightarrow \infty } m(h)= m(h)\).

Now we prove that \(\tilde{m}\) is idempotent. If \(g,g'\in \bar{C}(X,\mathbb {R})\) and \(f_n\rightarrow g,\,f'_n\rightarrow g'\), we have

$$\begin{aligned} \tilde{m}(g\oplus g')= & {} \lim _{n \rightarrow \infty } m(f_n\oplus f'_n) = \lim _{n \rightarrow \infty } [m(f_n)\oplus m(f'_n)] =\\= & {} [\lim _{n \rightarrow \infty } m(f_n)]\oplus [\lim _{n \rightarrow \infty } m(f'_n)]=\tilde{m}(g)\oplus \tilde{m}(g'). \end{aligned}$$

and \(\tilde{m}(a\odot g) = \lim _{n \rightarrow \infty } m(a\odot f_n) = \lim _{n \rightarrow \infty } [a\odot m(f_n)] = a\odot [\lim _{n \rightarrow \infty } m(f_n)]=a\odot \tilde{m}(g).\)

Finally, if C is a set closed with respect to \((\oplus ,\odot )\) operations, such that \(C(X,\mathbb {R})\subseteq C \subseteq \bar{C}(X,\mathbb {R})\) and \(\overline{m}\) is an idempotent measure on C which extends m, then given \(g\in C\) and any nonincreasing sequence of continuous function \((f_n)\) converging pointwise to g we have \(g\le f_n\). Therefore, \(\overline{m}(g) \le \overline{m}(f_n) = m(f_n )\) and taking the limit in n, we get \(\overline{m}(g) \le \tilde{m}(g)\). \(\square \)

Let us remember the definition of u.s.c. function.

Definition 6.9

A function \(f: X \rightarrow \mathbb {R}_{\max }\) is called upper semi-continuous (u.s.c. for short) at a point \(x_0 \in X\) if for every real \(c> f\left( x_0\right) \), there exists a neighborhood U of \(x_0\) such that \(f(x)< c\) for all \(x \in U\). Equivalently, f is upper semi-continuous at \(x_0\) if and only if

$$\begin{aligned} \limsup _{x \rightarrow x_0} f(x) \le f\left( x_0\right) . \end{aligned}$$

Next Lemma is inspired in [16, Lemma 1].

Lemma 6.10

Given an idempotent measure \(m \in C^{*}(X, \mathbb {R})\), consider the extension \(\tilde{m}\) from Lemma 6.8.

  1. 1.

    The map \(F:\mathbb {R} \times X \rightarrow \mathbb {R}_{\max }\) defined by \(F(a,x)=\tilde{m}(g^{a}_{x})\) is u.s.c. with respect to x and monotonous in \(a \in \mathbb {R}\).

  2. 2.

    \(F(a,x)=a \odot F(0,x)\), for any \(a \in \mathbb {R}, x \in X\);

  3. 3.
    $$\begin{aligned} m(h)=\bigoplus _{x \in X} F(h(x),x) = \bigoplus _{x \in X} F(0,x)\odot h(x), \end{aligned}$$

    for all \(h \in C(X, \mathbb {R})\).

Proof

(1) If \(a> a'\), then \(g_x^a \ge g_x^{a'}\). So \(F(a,x)= \tilde{m}(g_x^a) \ge \tilde{m}(g_x^{a'}) = F(a',x)\). We now prove that the correspondence \(x \rightarrow F(a,x)\) is u.s.c. Fix \(x_0 \in X\) and take any \(c > F(a,x_0)\). Recall that \(F(a,x_{0})=\tilde{m}(g^{a}_{x_{0}})= \lim _{n \rightarrow \infty } m(f_{n})< c\) for \((f_{n})\) the standard continuous approximation of \(g^{a}_{x_0}\). So there is \(N_{c}\) such that for all \(n_{0}> N_{c}\) we have, \(m(f_{n_{0}})< c\). Consider \(x \in U=B_{\frac{1}{n_{0}}}(x_{0})\) and choose n big enough so that \(f'_{n} \le f_{n_{0}}\) where \((f'_{n})\) is the standard continuous approximation of \(g^{a}_{x}\).

Under this hypothesis, we have \(m(f'_{n}) \le m(f_{n_{0}})\) and taking the limit on the left side, we get \(\lim _{n \rightarrow \infty }m(f'_{n}) \le m(f_{n_{0}})\), or equivalently \(F(a,x) \le m(f_{n_{0}})<c.\) This proves that the correspondence \(x \rightarrow F(a,x)\) is u.s.c.

(2) \(F(a,x)= \tilde{m}(g_x^a) = \tilde{m}(a\odot g_x^{0}) =a \odot \tilde{m}(g_x^{0})=a\odot F(0,x)\).

(3) By applying (2), we get the second equality and then we just need to prove that

$$\begin{aligned} m(h)=\bigoplus _{x \in X} F(h(x),x) \end{aligned}$$

for all \(h \in C(X, \mathbb {R})\). As \(g_x^{h(x)}(y) \le h(y),\, \forall \, x,y\in X\), we obtain \(\tilde{m}(g_x^{h(x)})\le \tilde{m}(h) = m(h)\,\forall \,x\in X\). Then

$$\begin{aligned} \bigoplus _{x \in X} F(h(x),x) = \bigoplus _{x \in X} \tilde{m}(g_x^{h(x)}) \le \bigoplus _{x \in X} m(h) = m(h). \end{aligned}$$

To prove the opposite inequality, we consider any \(\varepsilon >0\). Denote by \(f_n^x, n\ge 0\) the standard approximation of \(g_x^{h(x)}\) as given in (16). For each \(n\ge 0\), let \(x_n\in X\) be such that \(\displaystyle {\oplus _{x\in X}m(f_n^x) < m(f_n^{x_n})+\varepsilon }\). As X is compact, there exists a point \(\tilde{x}\) and a subsequence \((x_{n_i})\) such that \(x_{n_i}\rightarrow \tilde{x}\). As h is continuous, there exists \(k_0\) such that \(h(x)<h(\tilde{x})+\varepsilon \) for any \(x\in B(\tilde{x},\frac{1}{k_0}).\) For each natural \(k\ge k_0\), there exists \(N>k\) such that \(B(x_{n_i},\frac{2}{n_i})\subset B(\tilde{x},\frac{1}{k})\) for any \(n_i\ge N\). As \(h(x_{n_i})<h(\tilde{x})+\varepsilon \) and \(B(x_{n_i},\frac{2}{n_i})\subset B(\tilde{x},\frac{1}{k})\), it follows from definition (16) that \(f_{n_i}^{x_{n_i}} \le f_{k}^{\tilde{x}}+\varepsilon \) for \(n_i\ge N\) and consequently, for \(n_i\ge N \), we obtain

$$\begin{aligned} \oplus _{x\in X}m(f_{n_i}^x)< m(f_{n_i}^{x_{n_i}})+\varepsilon < m(f_{k}^{\tilde{x}})+2\varepsilon . \end{aligned}$$
(17)

As h is continuous and X is compact, the function h is uniformly continuous. Then there exists \(\delta \) such that \(d(x,y)<\delta \Rightarrow |h(y)-h(x)|<\varepsilon \). Consider a finite cover of X by balls of radius \(\frac{1}{n_i}< \delta \), with \(X\subset B(z_1,\frac{1}{n_i}) \cup \cdots \cup B(z_{l_i},\frac{1}{n_i})\). If \(y\in B(z_j,\frac{1}{n_i})\) then \(h(y) <h(z_j)+\varepsilon \). Furthermore, as from definition (16), \(f_{n_i}^{h(z_j)}(x)= h(z_j)\) for any \(x\in B(z_j,\frac{1}{n_i})\) we get, \(h \le \oplus _{z_j}f_{n_i}^{z_j} +\varepsilon .\) Finally, applying (17) we have,

$$\begin{aligned} m(h) \le m\left( \oplus _{z_j}f_{n_i}^{z_j} +\varepsilon \right) = \oplus _{z_j}m(f_{n_i}^{z_j})+\varepsilon \le \oplus _{x\in X}m(f_{n_i}^x)+\varepsilon \le m(f_{k}^{\tilde{x}})+3\varepsilon . \end{aligned}$$

Making \(k\rightarrow \infty \), we get

$$\begin{aligned} m(h) \le \tilde{m}(g_{\tilde{x}}^{h(\tilde{x})}) +3\varepsilon =F(h(\tilde{x}),\tilde{x}) +3\varepsilon \le \bigoplus _{x\in X}F(h(x),x) +3\varepsilon . \end{aligned}$$

As \(\varepsilon \) is arbitrary, we conclude the proof. \(\square \)

The next theorem corresponds to Theorem 1 of Ref. [16] in the present setting. It contains the claim in Theorem 1.2.

Theorem 6.11

For each \(\lambda \in U(X, \mathbb {R}_{\max })\), consider the functional \(m_{\lambda }: C(X,\mathbb {R}) \rightarrow \mathbb {R}_{\max }\) defined by

$$\begin{aligned} m_{\lambda }(h):=\bigoplus _{x \in X} \lambda (x) \odot h(x), \end{aligned}$$

for any \(h \in C(X,\mathbb {R})\). Then,

  1. 1.

    The map \(\gamma :U(X, \mathbb {R}_{\max }) \rightarrow C^{*}(X,\mathbb {R})\) defined by \(\gamma (\lambda )= m_{\lambda }\) is a max-plus isomorphism between \(U(X, \mathbb {R}_{\max })\) and \(C^{*}(X,\mathbb {R})\);

  2. 2.

    The function \(\lambda \in U(X, \mathbb {R}_{\max })\) is always bounded from above by \(m{_\lambda }(0)\) and \(m_\lambda \in I(X)\) is equivalent to \(\max _{x \in X} \lambda (x)=0\).

Proof

(1) Given \(\lambda \in U(X, \mathbb {R}_{\max }) \), let us prove that \(m_\lambda \in C^{*}(X,\mathbb {R})\). We have

$$\begin{aligned} m_\lambda (a\odot f) = \bigoplus _{x \in X} \lambda (x) \odot a\odot f(x) = a\odot \bigoplus _{x \in X} \lambda (x) \odot f(x) = a\odot m_\lambda ( f) \end{aligned}$$

and

$$\begin{aligned} m_\lambda (f\oplus g)= & {} \bigoplus _{x \in X} \lambda (x) \odot (f(x)\oplus g(x)) \\ {}= & {} \left( \bigoplus _{x \in X} \lambda (x) \odot f(x)\right) \oplus \left( \bigoplus _{x \in X} \lambda (x) \odot g(x)\right) = m_\lambda ( f)\oplus m_\lambda (g). \end{aligned}$$

We claim that \(\gamma \) is surjective. From Lemma 6.10, we know that for any \(m\in C^{*}(X,\mathbb {R}) \) we have

$$\begin{aligned} m(h) = \bigoplus _{x \in X} F(h(x),x )= \bigoplus _{x \in X} F(0,x) \odot h(x). \end{aligned}$$

Defining \(\lambda (x):= F(0,x)\), which is u.s.c. from Lemma  6.10, we obtain \(m=m_{\lambda }\).

We claim that \(\gamma \) is injective. Given \(\lambda \in U(X, \mathbb {R}_{\max })\) and \(m=m_\lambda \), let F(0, x) as given above. We will prove that \(\lambda (x) = F(0,x)\). Indeed, for a fixed \(x_0\), consider the standard continuous approximation \((f_n)\) of \(g^{0}_{x_0}\). As \(\lambda \) is u.s.c., we have

$$\begin{aligned} \lambda (x_0) = \lim _{n\rightarrow \infty } \bigoplus _{x\in X} \lambda (x)\odot f_n(x) \end{aligned}$$

Then

$$\begin{aligned} \lambda (x_0) = \lim _{n\rightarrow \infty } \bigoplus _{x\in X} \lambda (x)\odot f_n(x) = \lim _{n\rightarrow \infty } m_\lambda ( f_n) =\tilde{m}(g^{0}_{x_0}) = F(0,x_0). \end{aligned}$$

We need also to show that \(\gamma \) is max-plus linear. To see that take,

$$\begin{aligned} \gamma (a \odot \lambda )(h)= \bigoplus _{x \in X} a \odot \lambda (x) \odot h(x)= a \odot \bigoplus _{x \in X}\lambda (x) \odot h(x)= a \odot \gamma (\lambda )(h), \end{aligned}$$

and

$$\begin{aligned} \gamma (\lambda \oplus \lambda ')(h)= & {} \bigoplus _{x \in X} (\lambda (x) \oplus \lambda '(x)) \odot h(x)=\\= & {} \left( \bigoplus _{x \in X} \lambda (x) \odot h(x) \right) \oplus \left( \bigoplus _{x \in X} \lambda '(x) \odot h(x)\right) =\gamma (\lambda )(h) \oplus \gamma (\lambda ')(f), \end{aligned}$$

for all \(h \in C(X, \mathbb {R})\).

(2) We notice that \(m_\lambda (0) = \bigoplus _{x\in X} \lambda (x)\odot 0(x) = \bigoplus _{x\in X} \lambda (x)\), thus \(\lambda \) is always bounded by \(m_\lambda (0)\). Furthermore \(\bigoplus _{x \in X} \lambda (x)=0\) if and only if \(m_{\lambda }(0) = 0\). \(\square \)

Remark 6.12

The formula

$$\begin{aligned} m_{\lambda }(h):=\bigoplus _{x \in X} \lambda (x) \odot h(x), \end{aligned}$$

for any \(h \in C(X,\mathbb {R})\), in Theorem 6.11 can be slightly improved by recalling that, for each \(x \in X\) one can define the Dirac delta measure \(\delta _{x}: C(X,\mathbb {R}) \rightarrow \mathbb {R}\) by the formula \(\delta _{x}(h):= h(x).\) It is obvious that \(\delta _{x} \in C^*(X,\mathbb {R})\) (the density of \(\delta _{x}\) is \(g^{0}_{x}\)). Thus, we can see the previous formula as a kind of Choquet’s theorem (see [29, Theorem Choquet, Pg. 14]), where any idempotent measure is a max-plus combination of Dirac delta measures \(\delta _{x}\) with coefficients \(\lambda (x)\), that is, \(m_{\lambda }(h)=\bigoplus _{x \in X} \lambda (x) \odot \delta _{x}(h),\) for any \(f \in C(X,\mathbb {R})\) or,

$$\begin{aligned} m_{\lambda }=\bigoplus _{x \in X} \lambda (x) \odot \delta _{x}. \end{aligned}$$

1.2 Upper semi-continuous envelope

Definition 6.13

Let \(\mathcal {B}(X, \mathbb {R}_{\max })\) be the set of bounded functions in the sense of \(\rho \), that is, \(f: X \rightarrow \mathbb {R}_{\max }\) is bounded if there exists \(K>0\) such that \(d_{\rho }(f, -\infty )=\max _{x \in X}\rho (f(x), -\infty ) \le K\).

We notice that f is bounded with respect to \(\rho \) if, and only if, it is bounded from above in the usual sense because \(\rho (f(x), -\infty ) \le K \Leftrightarrow \exp (f(x))\le K \Leftrightarrow f(x) \le \ln (K).\) As X is compact any u.s.c. function (see Definition 6.9) is in \(\mathcal {B}(X, \mathbb {R}_{\max })\).

Definition 6.14

Given a function \(\lambda \in \mathcal {B}(X, \mathbb {R}_{\max })\), we define the upper semi-continuous envelope of \(\lambda \) as

$$\begin{aligned} \lambda ^{u.s.e.}(x)=\inf _{\varphi \in C(X, \mathbb {R}), \; \varphi \ge \lambda } \varphi (x), \; \forall x \in X. \end{aligned}$$

In Ref. [16], the lower semi-continuous envelope is used . The main properties of the upper semi-continuous envelope are given in the next lemma.

Proposition 6.15

By considering Definition 6.14,

  1. 1.

    If \(\lambda \in \mathcal {B}(X, \mathbb {R}_{\max })\) and \(\lambda \ne -\infty \) then \(\lambda ^{u.s.e.} \in U(X, \mathbb {R}_{\max })\);

  2. 2.

    If \(\lambda \in U(X, \mathbb {R}_{\max })\) then \(\lambda ^{u.s.e.}=\lambda \).

Proof

(1) First, we notice that \(\lambda \ne -\infty \) means that there exists \(x_0 \in X\) such that \(-\infty <\lambda (x_0)\le \lambda ^{u.s.e.}(x_0)\) thus, \(\textrm{supp}(\lambda ^{u.s.e.}) \ne \varnothing \). We claim that \(\lambda ^{u.s.e.}\) is u.s.c. Indeed, given \(x\in X\) and \(\varepsilon >0\), let \((x_n)\) be any sequence such that \(x_{n}\rightarrow x\). We fix any function \(\varphi _0 \in C(X, \mathbb {R})\) such that \(\varphi _0 \ge \lambda \). Then we have

$$\begin{aligned} \lambda ^{u.s.e.}(x_n)=\inf _{\varphi \in C(X, \mathbb {R}), \; \varphi \ge \lambda } \varphi (x_n) \le \varphi _0(x_n) \le \varphi _0(x)+\varepsilon \end{aligned}$$

for n big enough. Thus,

$$\begin{aligned} \limsup _{n\rightarrow \infty }\lambda ^{u.s.e.}(x_n) \le \varphi _0(x)+\varepsilon . \end{aligned}$$

If we take the infimum over functions \(\varphi _0\) at the right hand side of this inequality, we get

$$\begin{aligned} \limsup _{n\rightarrow \infty }\lambda ^{u.s.e.}(x_n) \le \lambda ^{u.s.e.}(x)+\varepsilon . \end{aligned}$$

As \(\varepsilon \) is arbitrary, we conclude that \(\lambda ^{u.s.e.}\) is u.s.c.. As \(\textrm{supp}( \lambda ^{u.s.e.}) \ne \varnothing \) we have proved that \(\lambda ^{u.s.e.} \in U(X, \mathbb {R}_{\max })\).

(2) As X is compact and \(\lambda \) is u.s.c., we know that \(\lambda \) attain its maximum value, which we will denote by M. Let \(x_0\) be a point of X. As \(\lambda \) is u.s.c. at \(x_0\), for any real c satisfying \(1+M>c> \lambda \left( x_0\right) \), there exists \(n\in \mathbb {N}\) such that \(\lambda (x)< c\) for all \(x \in B_{\frac{1}{n}}(x_0)\). Consider the continuous function \(\varphi _{c}\) defined by

$$\begin{aligned} \varphi _{c}(y):= \left\{ \begin{array}{ll} 1+M, &{}\quad d(y,x_0)>1/n \\ c+n(1+M-c)d(y,x_0), &{}\quad d(y,x_0)\le 1/n \end{array} \right. , \end{aligned}$$
(18)

and observe that \(\varphi _{c} \ge \lambda \). Thus, \(\lambda (x_0) \le \lambda ^{u.s.e.} (x_0) \le \varphi _c(x_0) =c.\) As \(c>\lambda (x_0)\) is arbitrary, we conclude that \(\lambda ^{u.s.e.}(x_0)=\lambda (x_{0})\). \(\square \)

Proposition 6.16

For each \(\lambda \in \mathcal {B}(X, \mathbb {R}_{\max }) \), consider the functional \(m_{\lambda }: C(X,\mathbb {R}) \rightarrow \mathbb {R}_{\max }\) defined by

$$\begin{aligned} m_{\lambda }(h):=\bigoplus _{x \in X} \lambda (x) \odot h(x), \end{aligned}$$

for any \(h \in C(X,\mathbb {R})\). If \(\lambda _{1}, \lambda _{2} \in \mathcal {B}(X, \mathbb {R}_{\max })\) and \(m_{\lambda _{1}}= m_{\lambda _{2}},\) then \(\lambda _{1}^{u.s.e.}= \lambda _{2}^{u.s.e.}\).

Proof

Since \(m_{\lambda _{1}}= m_{\lambda _{2}}\), we have

$$\begin{aligned} \bigoplus _{x \in X} \lambda _{1}(x) \odot h(x) = \bigoplus _{x \in X} \lambda _{2}(x) \odot h(x), \end{aligned}$$

for any \(h \in C(X,\mathbb {R})\). Given \(\varphi \in C(X, \mathbb {R})\) we have \(\varphi \ge \lambda _{1}\) if and only if \(\varphi \ge \lambda _{2}\). Indeed, suppose \(\varphi \ge \lambda _1\) and take \(h=-\varphi \). Then we get

$$\begin{aligned} 0 \ge \bigoplus _{x \in X} \lambda _{1}(x) -\varphi (x) = \bigoplus _{x \in X} \lambda _{2}(x) -\varphi (x), \end{aligned}$$

meaning that \(\varphi \ge \lambda _{2}\). The reverse argument is analogous.

Finally, fixed \(x_0 \in X\), we have

$$\begin{aligned} \lambda _{1}^{u.s.e.}(x_0)=\inf _{\varphi \in C(X, \mathbb {R}), \; \varphi \ge \lambda _{1}} \varphi (x_0) =\inf _{\varphi \in C(X, \mathbb {R}), \; \varphi \ge \lambda _{2}} \varphi (x_0) = \lambda _{2}^{u.s.e.}(x_0). \end{aligned}$$

\(\square \)

1.3 Maximal extension of idempotent measures

In this section, we present a brief discussion of how the domain of an idempotent measure can be extended from \(C(X,\mathbb {R})\) to \(\mathcal {B}(X, \mathbb {R}_{\max })\).

Proposition 6.17

Let \(\bar{C}(X,\mathbb {R})\) and \(\tilde{m}\) as defined in Lemma 6.8 where \(m=m_\lambda \). If \(g \in \bar{C}(X,\mathbb {R})\) then

$$\begin{aligned} \tilde{m}(g) = \bigoplus _{x\in X} \lambda (x)\odot g(x). \end{aligned}$$

Proof

Let \(f_n\) be a nonincreasing sequence of continuous functions converging pointwise to g. Denote by

$$\begin{aligned} \tilde{m}(g) = \lim _{n\rightarrow \infty } \bigoplus _{x\in X} \lambda (x)\odot f_n(x) \end{aligned}$$

and

$$\begin{aligned} M_\lambda (g) = \bigoplus _{x\in X} \lambda (x)\odot g(x). \end{aligned}$$

We want to prove that \(\tilde{m}(g) = M_\lambda (g)\). As \(f_n\ge g\) we have

$$\begin{aligned} \tilde{m}(g) = \lim _{n\rightarrow \infty } \bigoplus _{x\in X} \lambda (x)\odot f_n(x) \ge \bigoplus _{x\in X} \lambda (x)\odot g(x)=M_\lambda (g). \end{aligned}$$

To prove the equality suppose initially \(M_\lambda (g)\ne -\infty \) and by contradiction suppose there exists an \(\varepsilon >0\) and a sequence \((x_n)\) in X such that \(\lambda (x_n)\odot f_n(x_n) > M_\lambda (g)+\varepsilon . \) We can suppose there exists \(x_0\in X\) such that \(x_n\rightarrow x_0\).

Case 1: supposing \(\lambda (x_0)\odot g(x_0) \ne -\infty \). As \(\lambda (x_0)\odot f_n(x_0)\) converges to \(\lambda (x_0)\odot g(x_0)\), there exists k such that \(\lambda (x_0)\odot f_k(x_0)< \lambda (x_0)\odot g(x_0)+\varepsilon \le M_{\lambda }(g)+\varepsilon \). As \(\lambda +f_k\) is u.s.c., we get a \(\delta >0\) such that

$$\begin{aligned} d(x,x_0)<\delta \Rightarrow \lambda (x)\odot f_k(x) < M_\lambda (g)+\varepsilon . \end{aligned}$$

As \(f_n\) is nonincreasing, we get

$$\begin{aligned}{}[d(x,x_0)<\delta , n\ge k] \Rightarrow \lambda (x)\odot f_n(x) < M_\lambda (g)+\varepsilon , \end{aligned}$$

which is a contradiction because \(d(x_n,x_0)<\delta \) for n large enough.

Case 2: supposing \(\lambda (x_0)\odot g(x_0) = -\infty \). Then for any natural N, there exists k such that \(\lambda (x_0)\odot f_k(x_0)< -N\). As \(\lambda +f_k\) is u.s.c., we get a \(\delta >0\) such that

$$\begin{aligned} d(x,x_0)<\delta \Rightarrow \lambda (x)\odot f_k(x) <-N. \end{aligned}$$

As \(f_n\) is nonincreasing, we get

$$\begin{aligned}{}[d(x,x_0)<\delta , n\ge k] \Rightarrow \lambda (x)\odot f_n(x) < -N, \end{aligned}$$

which is a contradiction because \(d(x_n,x_0)<\delta \) for n large enough.

Now we suppose \(M_\lambda (g)= -\infty \), which means \(\lambda (x)+g(x) = -\infty ,\;\forall \,x\in X\), and by contradiction, suppose there exists a real number L and a sequence \((x_n)\) in X such that \(\lambda (x_n)\odot f_n(x_n) > L. \) We can suppose there exists \(x_0\in X\) such that \(x_n\rightarrow x_0\). Clearly \(\lambda (x_0)\odot g(x_0) = -\infty \) and we get a contradiction arguing as in case 2 above. \(\square \)

From Theorem 6.11, we have a formula for an idempotent measure \(m_{\lambda }\), with density \(\lambda \), which can be extended to \( h \in \mathcal {B}(X, \mathbb {R}_{\max })\) by setting up

$$\begin{aligned} \hat{m}_{\lambda }(h) = \bigoplus _{x \in X} \lambda (x) \odot h(x), \end{aligned}$$

where the supremum is well defined because h and \(\lambda \) are bounded from above. The functional \(\hat{m}_{\lambda }\) is called the maximal extension of \(m_{\lambda }\) (see Refs. [16, Corollary 1], [2, Theorem 2.2], and [18, Pg. 36], for a minimal version). The precise meaning of this name is explained by Lemma 6.8 and by Remark 6.21.

Definition 6.18

Given \(\lambda \in \mathcal {B}(X, \mathbb {R}_{\max })\), we define the idempotent integral for \(m_{\lambda } \in C^{*}(X,\mathbb {R})\) by the formula

$$\begin{aligned} \int _{X} h(x)\textrm{d}m_{\lambda }(x):= \hat{m}_{\lambda }(h) = \bigoplus _{x \in X} \lambda (x) \odot h(x), \end{aligned}$$

for all \( h \in \mathcal {B}(X, \mathbb {R}_{\max })\). Obviously, if \(h \in C(X, \mathbb {R})\) then \(\int _{X} h(x) \textrm{d}m_{\lambda }(x)= m_{\lambda }(h)\).

Definition 6.19

Given any \(A \subset X\), we define the max-plus indicator function of A as the function

$$\begin{aligned} \chi _{A}(x):= \left\{ \begin{array}{ll} 0, &{}\quad x \in A \\ -\infty , &{}\quad x \not \in A \end{array} \right. . \end{aligned}$$

Obviously, \(\chi _{A} \in \mathcal {B}(X, \mathbb {R}_{\max })\), for any \(A \subset X\), so we can compute the integral of \(\chi _{A}\) with respect to an idempotent measure \(m_{\lambda }\) by applying the extension \(\hat{m}_{\lambda }\)

$$\begin{aligned} \int _{X} \chi _{A} d\mu := \hat{m}_{\lambda }(\chi _{A}) = \bigoplus _{x \in X} \lambda (x) \odot \chi _{A}(x) = \bigoplus _{x \in A} \lambda (x). \end{aligned}$$

We recall that \(2^X\) is the set of parts of a set X.

Definition 6.20

Consider \(m_{\lambda } \in C^{*}(X,\mathbb {R})\) with extension \(\hat{m}_{\lambda }\). The correspondence (we use the same \(m_{\lambda }\) as symbol if there is no risk of confusion) \(m_{\lambda }: 2^X \rightarrow \mathbb {R}\) defined by

$$\begin{aligned} m_{\lambda }(A):= \int _{X} \chi _{A}(x) \textrm{d}m_{\lambda }(x) = \bigoplus _{x \in A} \lambda (x), \end{aligned}$$

for all \(A \in 2^X \) is called the set idempotent measure (see Refs. [18, Corollary 1], [2, Definition 2.1] for cost measures, [1, Section 3] and [11, Definition 1]).

Remark 6.21

By item (3) of Theorem 6.11, if \(\lambda \in U(X, \mathbb {R}_{\max })\) and \(\lambda _1 \in \mathcal {B}(X, \mathbb {R}_{\max })\) satisfy \(m_{\lambda _{1}}= m_{\lambda }\) then \(\lambda = \lambda _{1}^{u.s.e.} \ge \lambda _{1}\). Thus,

$$\begin{aligned} m_{\lambda }(A)= \bigoplus _{x \in A} \lambda (x) \ge \bigoplus _{x \in A} \lambda _{1}(x)=m_{\lambda _{1}}(A), \end{aligned}$$

for all \(A \in 2^X\). So, the choice of the u.s.c. function \(\lambda \) produces an integral bigger or equal to any other choice \(\lambda _1\), which explains the meaning of the name “maximal”.

Proposition 6.22

Let \(m_{\lambda } \in C^{*}(X,\mathbb {R})\) with extension \(\hat{m}_{\lambda }\). Then,

  1. 1.

    \(m_{\lambda }(\varnothing )=-\infty \);

  2. 2.

    \(m_{\lambda }(A \cup B)= m_{\lambda }(A ) \oplus m_{\lambda }(B)\) for any \(A,B \in 2^X\).

Proof

(1) As usual in the set function theory, we assume that, for any function \(f: X \rightarrow \mathbb {R}_{\max }\) the supremum over an empty set is the smallest value in \(\mathbb {R}_{\max }\), that is, \(-\infty \). Thus, \(m_{\lambda }(\varnothing )=\bigoplus _{x \in \varnothing } \lambda (x) = -\infty \).

(2)

$$\begin{aligned} m_{\lambda }(A \cup B)= & {} \int _{X} \chi _{_{A \cup B}}(x) \textrm{d}m_{\lambda }(x)= \hat{m}_{\lambda }(\chi _{_{A \cup B}}) = \bigoplus _{x \in X} \lambda (x) \odot \chi _{_{A \cup B}}(x) \\= & {} \bigoplus _{x \in X} \lambda (x) \odot ( \chi _{_{A}}(x) \oplus \chi _{_{B}}(x))\\{} & {} = \bigoplus _{x \in B} \lambda (x) \oplus \bigoplus _{x \in A} \lambda (x) = m_{\lambda }(A ) \oplus m_{\lambda }(B). \end{aligned}$$

\(\square \)

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

Mengue, J.K., Oliveira, E.R. Invariant measures for place-dependent idempotent iterated function systems. J. Fixed Point Theory Appl. 26, 19 (2024). https://doi.org/10.1007/s11784-024-01109-8

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11784-024-01109-8

Keywords

Mathematics Subject Classification

Navigation