Skip to main content

The Dual Potential, the Involution Kernel and Transport in Ergodic Optimization

  • Conference paper

Part of the book series: CIM Series in Mathematical Sciences ((CIMSMS,volume 1))

Abstract

Consider the shift σ acting on the Bernoulli space \(\varSigma =\{ 1,2,\ldots,n\}^{\mathbb{N}}\). We denote \(\hat{\varSigma }=\{ 1,2,\ldots,n\}^{\mathbb{Z}} =\varSigma \times \varSigma\). We analyze several properties of the maximizing probability μ , A of a Hölder potential \(A:\varSigma \rightarrow \mathbb{R}\). Associated to A(x), via the involution kernel, W(x, y), \(W:\hat{\varSigma }\rightarrow \mathbb{R}\), one can get the dual potential A (y), where \((x,y) \in \hat{\varSigma }\). We denote \(\mu _{\infty,A^{{\ast}}}\) the maximizing probability for A . We would like to consider the transport problem from μ , A to \(\mu _{\infty,A^{{\ast}}}\). In this case, it is natural to consider the cost function c(x, y) = I(x) − W(x, y) +γ, where I is the deviation function for μ , A , as the limit of Gibbs probabilities μ β A for the potential β A when β → . The value γ is a constant which depends on A. We could also take c = −W above. We denote by \(\mathcal{K} = \mathcal{K}(\mu _{\infty,A},\mu _{\infty,A^{{\ast}}})\) the set of probabilities \(\hat{\eta }(x,y)\) on \(\hat{\varSigma }\), such that \(\pi _{x}^{{\ast}}(\hat{\eta }) =\mu _{\infty,A},\,\,\text{and}\,\,\pi _{y}^{{\ast}}(\hat{\eta }) =\mu _{\infty,A^{{\ast}}}\,.\) We describe the minimal solution \(\hat{\mu }\) (which is invariant by the shift on \(\hat{\varSigma }\)) of the Transport Problem, that is, the solution of

$$\displaystyle{\inf _{\hat{\eta }\in \mathcal{K}}\int \int c(x,y)\,d\,\hat{\eta } =\, -\,\max _{\hat{\eta }\in \mathcal{K}}\int \int (W(x,y)-\gamma )\,d\,\hat{\eta }.\,}$$

The optimal pair of functions for the Kantorovich Transport dual Problem is (−V, −V ), where we denote the two calibrated sub-actions by V and V , respectively, for A and A . We show that the involution kernel W is cyclically monotone. In other words, satisfies a twist condition in the support of \(\hat{\mu }\). We analyze the question: is the support of \(\hat{\mu }\) a graph? We also investigate the question of finding an explicit expression for the function \(f:\varSigma \rightarrow \mathbb{R}\) whose c-subderivative determines the graph. We also analyze the same kind of problem for expanding transformations on the circle.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

References

  1. Bangert, V.: Mather sets for twist maps and geodesics on tori. In: Dynamics Reported, vol. 1, pp. 1–56. Wiley, Chichester (1988)

    Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  3. Baraviera, A.T., Cioletti, L.M., Lopes, A.O., Mohr, J., Souza, R.R.: On the general XY Model: positive and zero temperature, selection and non-selection. Rev. Math. Phys. 23(10), 1063–1113 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  4. Baraviera, A., Lopes, A.O., Mengue, J.: On the selection of subaction and measure for a subclass of potentials defined by P. Walters. Ergodic Theory Dyn. Syst. 33(5), 1338–1362 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  5. Baraviera, A., Leplaideur, R., Lopes, A.O.: Ergodic optimization, zero temperature limits and the max-plus algebra. In: Mini-course in XXIX Coloquio Brasileiro de Matemática, Rio de Janeiro (2013)

    MATH  Google Scholar 

  6. Bhattacharya, P., Majumdar, M.: Random Dynamical Systems. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  7. Bissacot, R., Garibaldi, E.: Weak KAM methods and ergodic optimal problems for countable Markov shifts. Bull. Braz. Math. Soc. 41(3), 321–338 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bousch, T.: Le poisson n’a pas d’arêtes. Ann. Inst. Henri Poincaré Probab. Stat. 36, 489–508 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  9. Bousch, T.: La condition de Walters. Ann. Sci. l’École Normale Supérieure 34, 287–311 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  10. Contreras, G.: Ground states are generically a periodic orbit, Arxiv (2013)

    Google Scholar 

  11. Contreras, G., Iturriaga, R.: Global minimizers of autonomous Lagrangians. 22 Colóquio Brasileiro de Matemática, IMPA (1999)

    Google Scholar 

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

    Google Scholar 

  13. Contreras, G., Lopes, A.O., Oliveira, E.: Ergodic Transport Theory, periodic maximizing probabilities and the twist condition. In: D. Zilberman, A. Pinto (eds.) Modeling, Optimization, Dynamics and Bioeconomy. Springer Proceedings in Mathematics, pp. 183–219. Springer, Cham (2014)

    Google Scholar 

  14. Conze, J.P., Guivarc’h, Y.: Croissance des sommes ergodiques et principe variationnel, manuscript circa (1993)

    Google Scholar 

  15. Delon, J., Salomon, J., Sobolevski, A.: Fast transport optimization for Monge costs on the circle. SIAM J. Appl. Math. 7, 2239–2258 (2010)

    Article  MathSciNet  Google Scholar 

  16. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications. Springer, New York (1998)

    Book  MATH  Google Scholar 

  17. Evans, L., Gomes, D.: Linear programming interpretation of Mather’s variational principle. ESAIM Control Optim. Cal. Var. 8, 693–702 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  18. Galatolo, S., Pacifico, M.: Lorenz-like flows: exponential decay of correlations for the Poincaré map, logarithm law, quantitative recurrence. Ergodic Theory Dyn. Syst. 30(6), 1703–1737 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  19. Gangbo, W., McCann, R.J.: The geometry of optimal transportation. Acta Math. 177, 113–161 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  20. Garibaldi, E., Lopes, A.O.: Functions for relative maximization. Dyn. Syst. 22, 511–528 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  21. Garibaldi, E., Lopes, A.O., Thieullen, Ph.: On calibrated and separating sub-actions. Bull. Braz. Math. Soc. 40(4), 577–602 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  22. Garibaldi, E., Lopes, A.O.: The effective potential and transshipment in thermodynamic formalism at temperature zero. Stoch. Dyn. 13(1), 1250009 (13 p) (2013)

    Google Scholar 

  23. Garibaldi, E., Thieullen, Ph.: Minimizing orbits in the discrete Aubry-Mather model. Nonlinearity 24(2), 563–611 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  24. Garibaldi, E., Thieullen, Ph.: Description of some ground states by Puiseux technics. J. Stat. 146(1), 125–180 (2012)

    MathSciNet  MATH  Google Scholar 

  25. Gole, C.: Sympletic Super-Twist Maps. World Scientific, Singapore (1998)

    Google Scholar 

  26. Hunt, B.R., Yuan, G.C.: Optimal orbits of hyperbolic systems. Nonlinearity 12, 1207–1224 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  27. Jenkinson, O.: Ergodic optimization. Discrete Continuous Dyn. Syst. Ser. A 15, 197–224 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  28. Jenkinson, O.: Every ergodic measure is uniquely maximizing. Discrete Continuous Dyn. Syst. Ser. A 16, 383–392 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Jenkinson, O.: A partial order on x2 -invariant measures. Math. Res. Lett. 15(5), 893–900 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  30. Jenkinson, O.: Optimization and majorization of invariant measures. Electron. Res. Announc. Am. Math. Soc. 13, 1–12 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  31. Jenkinson, O., Steel, J.: Majorization of invariant measures for orientation-reversing maps. Ergodic Theory Dyn. Syst. 30(5), 1471–1483 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  32. Kloeckner, B.: Optimal Transport and dynamics of circle expanding maps acting on measures. Ergodic Theory Dyn. Syst. 33(2), 529–548 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  33. Kloeckner, B., Lopes, A.O., Stadlbauer, M.: Contraction in the Wasserstein metric for some Markov chains, and applications to the dynamics of expanding maps, preprint (2014)

    Google Scholar 

  34. Kloeckner, B., Giulietti, P., Lopes, A.O., Marcon, D.: On the Geometry of Thermodynamical Formalism, preprint (2014)

    Google Scholar 

  35. Leplaideur, R.: A dynamical proof for the convergence of Gibbs measures at temperature zero. Nonlinearity 18(6), 2847–2880 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  36. Lopes, A., Mengue, J.: Duality theorems in ergodic transport. J. Stat. Phys. 149(5), 921–942 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  37. Lopes, A.O., Oliveira, E.R.: On the thin boundary of the fat attractor, preprint UFRGS (2011)

    Google Scholar 

  38. Lopes, A.O., Thieullen, Ph.: Sub-actions for Anosov diffeomorphisms. Astérisque 287, 135–146 (2003)

    MathSciNet  Google Scholar 

  39. Lopes, A.O., Thieullen, P.: Mather measures and the Bowen-Series transformation. Ann. Inst. Henri Poincaré Analyse non Linéaire 23, 663–682 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  40. Lopes, A.O., Oliveira, E.R., Smania, D.: Ergodic transport theory and piecewise analytic subactions for analytic dynamics. Bull. Braz. Math. Soc. 43(3), 467–512 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  41. Lopes, A.O., Mengue, J.K., Mohr, J., Souza, R.R.: Entropy, pressure and duality for Gibbs plans in ergodic transport. Bull. Braz. Math. Soc. (to appear)

    Google Scholar 

  42. Lopes, A., Mengue, J.K., Mohr, J., Souza, R.R.: Entropy and variational principle for one-dimensional lattice systems with a general a-priori probability: positive and zero temperature. Ergodic Theory Dyn. Syst. (to appear)

    Google Scholar 

  43. Mather, J.: Action minimizing invariant measures for positive definite Lagrangian Systems. Math. Z. 207(2), 169–207 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  44. Mengue, J.K., Oliveira, E.R.: Duality results for Iterated Function Systems with a general family of branches, preprint Arxiv (2014)

    Google Scholar 

  45. Mitra, T.: Introduction to dynamic optimization theory. In: Majumdar, M., Mitra, T., Nishimura, K. (eds.) Optimization and Chaos. Studies in Economic Theory. Springer, Heidelberg (2000)

    Google Scholar 

  46. Morris, I.D.: A sufficient condition for the subordination principle in ergodic optimization. Bull. Lond. Math. Soc. 39(2), 214–220 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  47. Parry, W., Pollicott, M.: Zeta functions and the periodic orbit structure of hyperbolic dynamics. In: Astérisque vol. 187–188. Société mathématique de France, Paris (1990)

    Google Scholar 

  48. Rachev, S., Ruschendorf, L.: Mass Transportation Problems, vol. 1 and 2. Springer, New York (1998)

    MATH  Google Scholar 

  49. Ruscheendorf, L.: On c-optimal random variables. Stat. Probab. Lett. 27, 267–270 (1996)

    Article  Google Scholar 

  50. Savchenko, S.V.: Cohomological inequalities for finite Markov chains. Funct. Anal. Appl. 33, 236–238 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  51. Souza, R.R.: Ergodic and Thermodynamic Games, preprint (to appear in Stochastics and Dynamics, 2014)

    Google Scholar 

  52. Tal, F.A., Zanata, S.A.: Maximizing measures for endomorphisms of the circle. Nonlinearity 21, 2347–2359 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  53. Villani, C.: Topics in Optimal Transportation. AMS, Providence (2003)

    Book  MATH  Google Scholar 

  54. Villani, C.: Optimal Transport: Old and New. Springer, Berlin (2009)

    Book  Google Scholar 

Download references

Acknowledgements

The first author is partially supported by CNPq, CAPES and FAPERGS.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. R. Oliveira .

Editor information

Editors and Affiliations

Appendix

Appendix

Here we consider first the shift \(\varSigma =\{ 0,1\}^{\mathbb{N}}\), and Σ as a metric space with the usual distance:

$$\displaystyle{d(x,y) = \left \{\begin{array}{cc} 0,\,\,\text{if}\, & \quad x = y\\ (1/2)^{n },\, &\,\,\text{if} \,\quad n =\min \{ i\,\,\vert \,\,x_{ i}\neq y_{i}\}. \end{array} \right.}$$

Additionally, we suppose that Σ is ordered by x < y, if x i  = y i for i = 1. . n − 1, and x n  = 0 and y n  = 1.

As the usual, we consider the dynamical system (Σ, σ) where σ: Σ → Σ is given by σ(x) = σ(x 1, x 2, x 3, ) = (x 2, x 3, x 4, ).

  1. (a)

    Potentials and the Involution Kernel

    As usual we denote

    $$\displaystyle{\tau _{x}^{{\ast}}(y) = (x_{ 1},y_{1},y_{2},y_{3},\ldots )\text{ and }\tau _{y}(x) = (y_{1},x_{1},x_{2},x_{3},\ldots ),}$$

    and

    $$\displaystyle{\hat{\sigma }(x,y) = (\sigma (x),\tau _{x}^{{\ast}}(y))\text{ and }\hat{\sigma }^{-1}(x,y) = (\tau _{ y}x,\sigma ^{{\ast}}(y)),}$$

    the skew product map, where σ (y = (y 1, y 2, y 3, )) = (y 2, y 3, y 4, ).

We also define τ k, y x = (y k , y k−1, … y 2, y 1, x 0, x 1, x 2, ), where x = (x 0, x 1, x 2, ),  y = (y 1, y 2, y 3, ). In a similar way we define τ k, y x. 

Given a continuous function \(A:\varSigma \rightarrow \mathbb{R}\), remember that a continuous function \(W:\varSigma \times \varSigma \rightarrow \mathbb{R}\) is an involution kernel for A if \((W \circ \hat{\sigma }^{-1} - W + A \circ \hat{\sigma }^{-1})(x,y)\) does not depends on x; In this case the continuous function \(A^{{\ast}}(y) = (W \circ \hat{\sigma }^{-1} - W + A \circ \hat{\sigma }^{-1})(x,y)\) is called the W-dual potential of A.

As in [2] we define the cocycle Δ A (x, x′, y), where

$$\displaystyle{\varDelta _{A}(x,x',y) =\sum _{n\geq 1}A \circ \hat{\sigma }^{-n}(x,y) - A \circ \hat{\sigma }^{-n}(x',y) =\sum _{ n\geq 1}A \circ \tau _{n,y}(x) - A \circ \tau _{n,y}(x'),}$$

and its dual version \(\varDelta _{A^{{\ast}}}(x,y,y')\), where

$$\displaystyle{\varDelta _{A^{{\ast}}}(x,y,y') =\sum _{n\geq 1}A^{{\ast}}\circ \hat{\sigma }^{n}(x,y) - A^{{\ast}}\circ \hat{\sigma }^{n}(x,y') =\sum _{ n\geq 1}A^{{\ast}}\circ \tau _{ n,x}^{{\ast}}(y) - A^{{\ast}}\circ \tau _{ n,x}^{{\ast}}(y').}$$

Note that:

  1. (i)

    Δ A (x, x′, y) = −Δ A (x′, x, y), in particular Δ A (x, x, y) = 0,

  2. (ii)

    Δ A (x, x′, y) +Δ A (x′, x″, y) = Δ A (x, x″, y),

  3. (iii)

    Δ A (x, x′, y) = Δ A (τ y x, τ y x′, σ (y)) + [Aτ y xAτ y x′], 

and the same relations are true for \(\varDelta _{A^{{\ast}}}(x,y,y')\).

Using this properties one can prove that, for any involution kernel we have \(W(x,y) - W(x',y) =\varDelta _{A}(x,x',y)\text{ and }W(x,y) - W(x,y') =\varDelta _{A^{{\ast}}}(x,y,y').\)

From this fact, we get that the difference between two involution kernels for A is a continuous function of y: {Involution kernels for A}∕C0(Σ) = W 0, where W 0(x, y) = Δ A (x, x′, y) for a fix x′ ∈ Σ is called a fundamental involution kernel of A. Indeed, the property (iii) shows that W 0 is an involution kernel for A.

On the other hand, given another involution kernel, W we have W(x, y) − W(x′, y) = Δ A (x, x′, y), thus

$$\displaystyle{W(x,y) = W(x',y) +\varDelta _{A}(x,x',y) = W(x',y) + W^{0}(x,y) = g(y) + W^{0}(x,y),}$$

where g(y) = W(x′, y) ∈ C 0(Σ).

As an example we compute the general dual potential. First for W 0(x, y) = Δ A (x, x′, y) we get:

$$\displaystyle{\begin{array}{ccc} A_{0}^{{\ast}}(y)& =&(W^{0}(\tau _{y}x,\sigma ^{{\ast}}(y))) - W^{0}(x,y) + A(\tau _{y}x) \\ & =& \varDelta _{A}(\tau _{y}x,x',\sigma ^{{\ast}}(y)) -\varDelta _{A}(x,x',y) + A(\tau _{y}x) \\ & =& A(\tau _{y}x') +\varDelta _{A}(\tau _{y}x',x',\sigma ^{{\ast}}(y)).\\ \end{array} }$$

Given another involution kernel, W we have W(x, y) = W(x′, y) + W 0(x, y) thus

$$\displaystyle{A^{{\ast}}(y) = (W \circ \hat{\sigma }^{-1} - W + A \circ \hat{\sigma }^{-1})(x,y) = W(x',\sigma ^{{\ast}}(y)) - W(x',y) + A_{ 0}^{{\ast}}(y).}$$
  1. (b)

    The Twist Property of an Involution Kernel

If \(A:\varSigma \rightarrow \mathbb{R}\) is a potential and W an arbitrary involution kernel for A, as we said before, W has the twist property, if for any, a, b, a′, b′ ∈ Σ

$$\displaystyle{W(a,b) + W(a',b') < W(a,b') + W(a',b),}$$

provided that a < a′ and b < b′.

If we rewrite this inequality as,

$$\displaystyle{\begin{array}{ccc} W(a,b) + W(a',b')& <&W(a,b') + W(a',b) \\ W(a,b) - W(a',b)& <&W(a,b') - W(a',b') \\ \varDelta _{A}(a,a',b) & <& \varDelta _{A}(a,a',b'), \end{array} }$$

we get an alternative criteria for the twist property, that is, W has the twist property, if for any, a, a′ ∈ Σ the function y → Δ A (a, a′, y), is strictly increasing, provided that a < a′.

Remark 5

This characterization shows a very important fact. The twist property is a property of A, so we can said that A is a twist potential or equivalently A has a twist involution kernel (as, obviously other involution kernel is also twist).

Remark 6

As an initial approximation we can consider a different setting of dynamics. Let T(x) = −2x ( mod 1 ), and

$$\displaystyle{\tau _{0}x = -\frac{1} {2}x + \frac{1} {2},\text{ and }\tau _{1}x = -\frac{1} {2}x + 1,}$$

the inverse branches that defines the skew maps (that are not the actual natural extension of T):

$$\displaystyle{\hat{T}(x,y) = (T(x),\tau _{x}^{{\ast}}(y))\text{ and }\hat{T}^{-1}(x,y) = (\tau _{ y}x,T^{{\ast}}(y)).}$$

So, one can compute an involutive (that is, A (y) = A(y)) smooth kernel for A 1(x) = x and A 2(x) = x 2 given by

$$\displaystyle{W_{1}(x,y) = -\frac{1} {3}(x + y)\text{ and }W_{2}(x,y) = \frac{1} {3}(x^{2} + y^{2}) -\frac{4} {3}xy.}$$

As a corollary we get that any potential A(x) = a + bx + cx 2 has a smooth involution kernel given by W(x, y) = a + bW 1(x, y) + cW 2(x, y). 

Here and in the next paragraphs, we will denote

$$\displaystyle{W_{A}(x,y):= a + bW_{1}(x,y) + cW_{2}(x,y),}$$

where A(x) = a + bx + cx 2 is a polynomial of degree 2.

We observe that the twist property can be derived from the positivity of the second mix derivative of the involution kernel when it is smooth. Note that,

$$\displaystyle{\frac{\partial ^{2}W_{1}} {\partial x\partial y} = 0,\text{ and }\frac{\partial ^{2}W_{2}} {\partial x\partial y} = -\frac{4} {3},}$$

thus W 1 is not twist and W 2 is. Actually any potential A(x) = a + bx + cx 2 where c > 0 is twist.

Remark 7

In this remark we are going to consider the case of A(x) = a + bx + cx 2 where c < 0 (not twist). In this case we will be able to compute the calibrated subaction explicitly, which, we believe, it is interesting in itself.

As a first example consider A(x) = −(x − 1)2 which is a convex potential.

From [30, 31] we get that the unique maximizing measure for this potential is μ  = δ 2∕3, so the critical value is m = A(2∕3). Using the fact that m = A(2∕3) one can show that there is a unique (up to constants) calibrated subaction ϕ given by:

$$\displaystyle{\phi (x) = W(x,2/3) - W(2/3,2/3) = -\frac{1} {3}x^{2} + \frac{2} {9}x}$$

where the kernel is given by

$$\displaystyle{W(x,y) = -(1/3)x^{2} - (1/3)y^{2} + (4/3)xy - (2/3)x - (2/3)y.}$$

As a second example consider \(A(x) = -(x -\frac{1} {2})^{2}\) which it is also a concave potential.

The general arguments in [31] shown that any maximizing measure for this potential is μ  = (1 − t)δ 1∕3 + t δ 2∕3, where t ∈ [0, 1], so the critical value is m = A(1∕3) = A(2∕3). In this case the involutive smooth involution kernel is:

$$\displaystyle{W(x,y) = -(1/3)x^{2} - (1/3)y^{2} + (4/3)xy - (2/3)x - (1/3)y.}$$

It is easy to verify that,

$$\displaystyle{\phi (x) = V _{1}(x)\chi _{_{[(0,1/2)]}}(x) + V _{2}(x)\chi _{_{[1/2,1]}}(x) =\max \{ V _{1}(x),V _{2}(x)\},}$$

is indeed a calibrated subaction for A, where

V 1(x) = W(x, 1∕3) − W(1∕3, 1∕3) = Δ(x, 1∕3, 1∕3) = −(1∕3)x 2 + (1∕9)x,

V 2(x) = W(x, 2∕3)−W(2∕3, 2∕3) = Δ(x, 2∕3, 2∕3) = −(1∕3)x 2+(5∕9)x−2∕9,

Note that,

$$\displaystyle{\begin{array}{ccc} \phi (\tau _{0}x)& =& V _{1}(\tau _{0}x)\chi _{_{[(0,1/2)]}}(\tau _{0}x) + V _{2}(\tau _{0}x)\chi _{_{[1/2,1]}}(\tau _{0}x) \\ & =& V _{1}(\tau _{0}x) =\varDelta (\tau _{0}x,1/3,1/3) \\ & =& \varDelta (\tau _{1/3}x,\tau _{1/3}1/3,T^{{\ast}}1/3) \\ & =&\varDelta (x,1/3,1/3) - [A(\tau _{1/3}x) - A(\tau _{1/3}1/3)] \\ & =& V _{1}(x) - [A(\tau _{0}x) - m].\\ \end{array} }$$

Thus ϕ(τ 0 x) + A(τ 0 x) − m = V 1(x). Analogously, ϕ(τ 1 x) + A(τ 1 x) − m = V 2(x) so

$$\displaystyle{\begin{array}{ccc} \phi (x)& =& \max \{V _{1}(x),V _{2}(x)\} \\ & =&\max \{\phi (\tau _{0}x) + A(\tau _{0}x) - m,\phi (\tau _{1}x) + A(\tau _{1}x) - m\} \\ & =& \max _{y\in \varSigma }\{\phi (\tau _{y}x) + A(\tau _{y}x) - m\}.\\ \end{array} }$$
  1. (c)

    Twist Criteria

Is natural to consider a criteria for the twist property for a class of functions that has a small dependence on the cubic (or higher order) terms. Let P 2 + = { p(x) = a + bx + cx 2 | c > 0} be the set of strictly convex polynomial. Consider p ∈ P 2 +, and define

$$\displaystyle{\mathcal{C}_{\varepsilon }(p) =\{ A \in \text{C}^{3}([0,1])\vert A(x) = p(x) +\varepsilon R(x),\,\text{where}\,\frac{\partial R} {\partial x} \in \text{C}^{3}([0,1])\}}$$

Theorem 8

For any p ∈ P 2 + , there exists \(\varepsilon > 0\) such that all \(A \in \mathcal{C}_{\varepsilon }(p)\) is twist.

Proof

Consider p ∈ P 2 + fixed. So, p has a smooth and involutive involution kernel given by

$$\displaystyle{W_{p}(x,y) = (a + bW_{1} + cW_{2})(x,y),}$$

that is, p (y) = p(y), where \(W_{1}(x,y) = -\frac{1} {3}(x + y)\) and \(W_{2}(x,y) = \frac{1} {3}(x^{2} + y^{2}) -\frac{4} {3}xy\), are the involution kernel associated to x and x 2 respectively. Let, \(A = p +\varepsilon R \in \mathcal{C}_{\varepsilon }(p)\), and W R be the involution kernel for R. Since R is C3 we get that, its corresponding involution kernel W R is C2 in the variable x. Using the linearity of the cohomological equation, we get \(W_{A}(x,y) = p(W)(x,y) +\varepsilon W_{R}(x,y)\), and differentiating with respect to x, we have

$$\displaystyle{\begin{array}{ccc} \frac{\partial } {\partial x}W_{A}(x,y)& =&(b \frac{\partial } {\partial x}W_{1} + c \frac{\partial } {\partial x}W_{2})(x,y) +\varepsilon \frac{\partial } {\partial x}W_{R}(x,y)\, = \\ & & -\frac{1} {3}b + \frac{2} {3}cx -\frac{4} {3}cy +\varepsilon \frac{\partial } {\partial x}W_{R}(x,y)\\ \end{array} }$$

Since \(-\frac{4} {3}c < 0\), and \(\frac{\partial } {\partial x}W_{R}(x,y) \in \text{C}^{0}([0,1]^{2})\) the compactness of [0, 1]2 implies that \(\frac{\partial } {\partial x}W_{A}(x,\cdot )\) is a strictly decreasing function for any \(\varepsilon\) small enough, which is sufficient to ensure the twist property.

Remark 8

If, A ∈ C([0, 1]) is strongly convex, we can consider a perturbation of A of order 2 given by

$$\displaystyle{B_{\varepsilon }(x) = A(0) - A'(0)x + \frac{A''(0)} {2} x^{2} +\varepsilon \sum _{ n\geq 3}\frac{A^{(n)}(0)} {n!} x^{n} \in \mathcal{C}_{\varepsilon }(p_{ A}),}$$

where \(p_{A} = A(0) - A'(0)x + \frac{A''(0)} {2} x^{2} \in P_{ 2}^{+}\). Thus, we can find \(\varepsilon _{0} > 0\) such that \(B_{\varepsilon }\) is twist for any \(0 <\varepsilon <\varepsilon _{0}\).

  1. (d)

    The Involution Kernel is Bi-Hölder

We consider now T(x) = 2x (mod 1) on the interval [0, 1] and the shift σ on \(\varOmega =\{ 0,1\}^{\mathbb{N}}\). A natural question is the regularity of the involution kernel W. We denote τ j , j = 0, 1 the two inverse branches of T. Given \(w = (w_{1},w_{2},\ldots ) \in \{ 0,1\}^{\mathbb{N}}\) we denote by τ k, w the transformation in [0, 1] given by \(\tau _{k,w}(x) = (\tau _{w_{k}} \circ \tau _{w_{k-1}} \circ \,\ldots \,\circ \tau _{w_{1}})\,(x).\) We have that, for a fixed x 0

$$\displaystyle{\varDelta (x,x_{0},w) =\sum _{ k=1}^{\infty }A(\tau _{ k,w}(x)) - A(\tau _{k,w}(x_{0}))}$$

and, the involution kernel W can be described as: for any (x, w) we have W(x, w) = Δ(x, x 0, w). It is easy to see that W is Hölder on the variable x. Consider a, b ∈ Ω and suppose that d(a, b) = 2n. In this way a j  = b j , j = 1, 2, n − 1, n. We denote \(\bar{a} =\sigma ^{n}(a)\) and \(\bar{b} =\sigma ^{n}(b)\).

Proposition 7

Suppose A is α−Hölder. Consider a,b ∈Ω such that d(a,b) = 2 −n . For a fixed x ∈ [0,1] we have | W(x,a) − W(x,b) |≤ C (2 −n ) α .

Proof

Note that for z = τ n, a (x) = τ n, b (x) and z 0 = τ n, a (x 0) = τ n, b (x 0) we have

$$\displaystyle\begin{array}{rcl} & W(x,a) - W(x,b) =\sum _{ k=1}^{\infty }A(\tau _{k,a}(x)) - A(\tau _{k,a}(x_{0})) - A(\tau _{k,b}(x)) + A(\tau _{k,b}(x_{0})) =& {}\\ & \sum _{k=1}^{\infty }[\,A(\tau _{k,a}(x)) - A(\tau _{k,b}(x))\,] - [\,A(\tau _{k,a}(x_{0})) - A(\tau _{k,b}(x_{0}))\,] = & {}\\ & \sum _{k=1}^{\infty }[\,A(\tau _{k,\bar{a}}(z)) - A(\tau _{k,\bar{b}}(z))\,] - [\,A(\tau _{k,\bar{a}}(z_{0})) - A(\tau _{k,\bar{b}}(z_{0}))\,]. & {}\\ \end{array}$$

Note also that | zz 0 | ≤ d(a, b) = 2n. Consider z = z 0 + h, then

$$\displaystyle\begin{array}{rcl} & \,A(\tau _{k,\bar{a}}(z_{0} + h)) - A(\tau _{k,\bar{a}}(z_{0})) \leq C_{A}\,d(\tau _{k,\bar{a}}(z_{0} + h),\tau _{k,\bar{a}}(z_{0}))^{\alpha } \leq & {}\\ & C_{A}\,(\,2^{-k}\,h\,)^{\alpha } = C_{A}(\,2^{-k}\,)^{\alpha }\,h^{\alpha }. & {}\\ \end{array}$$

Then,

$$\displaystyle\begin{array}{rcl} & \sum _{k=1}^{\infty }[\,A(\tau _{k,\bar{a}}(z)) - A(\tau _{k,\bar{a}}(z_{0}))\,] - [\,A(\tau _{k,\bar{b}}(z)) - A(\tau _{k,\bar{b}}(z_{0}))\,]& {}\\ & \leq C_{A}\sum _{k=1}^{\infty }2\,(\,2^{-k}\,)^{\alpha }\,h^{\alpha } \leq C_{A}\sum _{k=1}^{\infty }2\,(2^{\alpha })^{-k}\,\,h^{\alpha } \leq C\,d(a,b)^{\alpha }. & {}\\ \end{array}$$

From the above we get:

Theorem 9

If \(A: S^{1} \rightarrow \mathbb{R}\) is Hölder then \(W: S^{1} \times \{ 0,1\}^{\mathbb{N}} \rightarrow \mathbb{R}\) is bi-Hölder.

  1. (e)

    The Fenchel-Rockafellar Theorem Given \(f: \mathbb{R} \rightarrow \mathbb{R}\) defined on the variable x, the Legendre transform of f, denoted by f , is the function on the variable p defined by

    $$\displaystyle{f^{{\ast}}(p) =\sup _{ x\in \mathbb{R}}\{p\,x -\, f(x)\}.}$$

Theorem 10 (Fenchel-Rockafellar)

Suppose f(x) is smooth strictly convex, \(f: \mathbb{R} \rightarrow \mathbb{R}\) , and, g(x) is smooth strictly concave, \(g: \mathbb{R} \rightarrow \mathbb{R}\) . Denote by f and g the corresponding Legendre transforms on the variable p. Then,

$$\displaystyle{\inf _{x\in \mathbb{R}}\,\{\,f(x)\, -\, g(x)\,\}\, =\,\sup _{p\in \mathbb{R}}\,\{\,g^{{\ast}}(p)\, -\, f^{{\ast}}(p)\,\}}$$

Proof

By convexity and concavity properties we have that there exists x 0 such that

$$\displaystyle{\inf _{x\in \mathbb{R}}\,\{\,f(x)\, -\, g(x)\,\} = f(x_{0}) - g(x_{0}).}$$

It is also true that f′(x 0) − g′(x 0) = 0. Denote by \(\overline{p}\) that value \(\overline{p} = f'(x_{0})\). We illustrate the proof via two pictures in a certain particular case. Figure 11 shows a geometric picture of the position and values of f(x 0) − g(x 0), \(g^{{\ast}}(\overline{p})\) and \(f^{{\ast}}(\overline{p}).\) Note that in this picture we have that f(x 0) − g(x 0) > 0. This picture also shows the graph of \(\overline{p}\,x\) as a function of x. We observe that the Legendre transform is not linear on the function. Let’s consider different values of p and estimate f (p) and g (p). Suppose first \(p\, >\, \overline{p}\). In Fig. 12 we show the graph of px, and the values of f (p) and g (p). We denote by x 2 the value such that

$$\displaystyle{f^{{\ast}}(p) =\sup _{ x\in \mathbb{R}}\{p\,x -\, f(x)\}\, =\, p\,x_{2} - f(x_{2}).}$$

Note that x 2  > x 0. We denote by x 1 the value such that

$$\displaystyle{0\, <\, g^{{\ast}}(p) =\sup _{ x\in \mathbb{R}}\{p\,x -\, g(x)\}\, =\, p\,x_{1} - g(x_{1}).}$$

Note that x 1  < x 0.

Fig. 11
figure 11

The infimum

Fig. 12
figure 12

The supremum

Note also that f (p) and g (p) have different signs. From this picture one can see that g (p) − f (p) < f(x 0) − g(x 0). In the case \(p\, <\, \overline{p}\) a similar reasoning can be done.

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this paper

Cite this paper

Lopes, A.O., Oliveira, E.R., Thieullen, P. (2015). The Dual Potential, the Involution Kernel and Transport in Ergodic Optimization. In: Bourguignon, JP., Jeltsch, R., Pinto, A., Viana, M. (eds) Dynamics, Games and Science. CIM Series in Mathematical Sciences, vol 1. Springer, Cham. https://doi.org/10.1007/978-3-319-16118-1_20

Download citation

Publish with us

Policies and ethics