Skip to main content
Log in

Geometrical Models for a Class of Reducible Pisot Substitutions

  • Published:
Discrete & Computational Geometry Aims and scope Submit manuscript

Abstract

We set up a geometrical theory for the study of the dynamics of reducible Pisot substitutions. It is based on certain Rauzy fractals generated by duals of higher dimensional extensions of substitutions. We obtain under certain hypotheses geometric representations of stepped surfaces and related polygonal tilings, as well as self-replicating and periodic tilings made of Rauzy fractals. We apply our theory to one-parameter family of substitutions. For this family, we analyze and interpret in a new combinatorial way the codings of a domain exchange defined on the associated fractal domains. We deduce that the symbolic dynamical systems associated with this family of substitutions behave dynamically as first returns of toral translations.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

References

  1. Akiyama, S., Barge, M., Berthé, V., Lee, J.-Y., Siegel, A.: On the Pisot substitution conjecture. In: Kellendonk, J., Lenz, D., Savinien, J. (eds.) Mathematics of Aperiodic Order. Progress in Mathematics, vol. 309, pp. 33–72. Birkhäuser, Basel (2015)

    Chapter  Google Scholar 

  2. Arnoux, P., Bernat, J., Bressaud, X.: Geometrical models for substitutions. Exp. Math. 20(1), 97–127 (2011)

    Article  MathSciNet  Google Scholar 

  3. Arnoux, P., Berthé, V., Fernique, T., Jamet, D.: Functional stepped surfaces, flips, and generalized substitutions. Theor. Comput. Sci. 380(3), 251–265 (2007)

    Article  MathSciNet  Google Scholar 

  4. Arnoux, P., Furukado, M., Harriss, E., Ito, S.: Algebraic numbers, free group automorphisms and substitutions on the plane. Trans. Am. Math. Soc. 363(9), 4651–4699 (2011)

    Article  MathSciNet  Google Scholar 

  5. Arnoux, P., Ito, S.: Pisot substitutions and Rauzy fractals. Bull. Belg. Math. Soc. Simon Stevin 8(2), 181–207 (2001)

    MathSciNet  MATH  Google Scholar 

  6. Baker, V., Barge, M., Kwapisz, J.: Geometric realization and coincidence for reducible non-unimodular Pisot tiling spaces with an application to \(\beta \)-shifts. Ann. Inst. Fourier (Grenoble) 56(7), 2213–2248 (2006)

    Article  MathSciNet  Google Scholar 

  7. Barge, M.: The Pisot conjecture for \(\beta \)-substitutions. arXiv:1505.04408. (2015)

  8. Barge, M.: Pure discrete spectrum for a class of one-dimensional substitution tiling systems. Discrete Contin. Dyn. Syst. 36(3), 1159–1173 (2016)

    Article  MathSciNet  Google Scholar 

  9. Barge, M., Bruin, H., Jones, L., Sadun, L.: Homological Pisot substitutions and exact regularity. Isr. J. Math. 188, 281–300 (2012)

    Article  MathSciNet  Google Scholar 

  10. Berthé, V., Bourdon, J., Jolivet, T., Siegel, A.: A combinatorial approach to products of Pisot substitutions. Ergod. Theory Dyn. Syst. 36(6), 1757–1794 (2016)

    Article  MathSciNet  Google Scholar 

  11. Berthé, V., Delecroix, V.: Beyond substitutive dynamical systems: \(S\)-adic expansions. In: Berthé, V., et al. (eds.) Numeration and Substitution 2012. RIMS Kôkyûroku Bessatsu, vol. B46, pp. 81–123. Research Institute for Mathematical Sciences, Kyoto (2014)

  12. Berthé, V., Fernique, T.: Brun expansions of stepped surfaces. Discrete Math. 311(7), 521–543 (2011)

    Article  MathSciNet  Google Scholar 

  13. Berthé, V., Siegel, A.: Tilings associated with beta-numeration and substitutions. Integers 5(3), Art. No. A2 (2005)

  14. Berthé, V., Siegel, A., Thuswaldner, J.: Substitutions, Rauzy fractals and tilings. In: Berthé, V., Rigo, M. (eds.) Combinatorics, Automata and Number Theory. Encyclopedia of Mathematics and Its Applications, vol. 135, pp. 248–323. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  15. Canterini, V., Siegel, A.: Geometric representation of substitutions of Pisot type. Trans. Am. Math. Soc. 353(12), 5121–5144 (2001)

    Article  MathSciNet  Google Scholar 

  16. Clark, A., Sadun, L.: When size matters: subshifts and their related tiling spaces. Ergod. Theory Dyn. Syst. 23(4), 1043–1057 (2003)

    Article  MathSciNet  Google Scholar 

  17. Dumont, J.-M., Thomas, A.: Systemes de numeration et fonctions fractales relatifs aux substitutions. Theor. Comput. Sci. 65(2), 153–169 (1989)

    Article  MathSciNet  Google Scholar 

  18. Ei, H., Ito, S.: Tilings from some non-irreducible, Pisot substitutions. Discrete Math. Theor. Comput. Sci. 7(1), 81–121 (2005)

    MathSciNet  MATH  Google Scholar 

  19. Ei, H., Ito, S., Rao, R.: Atomic surfaces, tilings and coincidences. II. Reducible case. Ann. Inst. Fourier (Grenoble) 56(7), 2285–2313 (2006)

    Article  Google Scholar 

  20. Enomoto, F., Ei, H., Furukado, M., Ito, S.: Tilings of a Riemann surface and cubic Pisot numbers. Hiroshima Math. J. 37(2), 181–210 (2007)

    MathSciNet  MATH  Google Scholar 

  21. Fernique, T.: Multidimensional Sturmian sequences and generalized substitutions. Int. J. Found. Comput. Sci. 17(3), 575–599 (2006)

    Article  MathSciNet  Google Scholar 

  22. Frank, N.P.: A primer of substitution tilings of the Euclidean plane. Expo. Math. 26(4), 295–326 (2008)

    Article  MathSciNet  Google Scholar 

  23. Frettlöh, D.: Duality of model sets generated by substitutions. Rev. Roumaine Math. Pures Appl. 50(5–6), 619–639 (2005)

    MathSciNet  MATH  Google Scholar 

  24. Frougny, C., Solomyak, B.: Finite beta-expansions. Ergod. Theory Dyn. Syst. 12(4), 713–723 (1992)

    Article  MathSciNet  Google Scholar 

  25. Furukado, M.: Tilings from non-Pisot unimodular matrices. Hiroshima Math. J. 36(2), 289–329 (2006)

    MathSciNet  MATH  Google Scholar 

  26. Furukado, M., Ito, S., Robinson Jr., E.A.: Tilings associated with non-Pisot matrices. Ann. Inst. Fourier (Grenoble) 56(7), 2391–2435 (2006)

    Article  MathSciNet  Google Scholar 

  27. Hatcher, A.: Algebraic Topology. Cambridge University Press, Cambridge (2002)

    MATH  Google Scholar 

  28. Ito, S., Ohtsuki, M.: Modified Jacobi–Perron algorithm and generating Markov partitions for special hyperbolic toral automorphisms. Tokyo J. Math. 16(2), 441–472 (1993)

    Article  MathSciNet  Google Scholar 

  29. Ito, S., Ohtsuki, M.: Parallelogram tilings and Jacobi–Perron algorithm. Tokyo J. Math. 17(1), 33–58 (1994)

    Article  MathSciNet  Google Scholar 

  30. Ito, S., Rao, H.: Atomic surfaces, tilings and coincidence. I. Irreducible case. Isr. J. Math. 153, 129–155 (2006)

    Article  MathSciNet  Google Scholar 

  31. Lagarias, J.C., Wang, Y.: Substitution Delone sets. Discrete Comput. Geom. 29(2), 175–209 (2003)

    Article  MathSciNet  Google Scholar 

  32. Minervino, M., Thuswaldner, J.: The geometry of non-unit Pisot substitutions. Ann. Inst. Fourier (Grenoble) 64(4), 1373–1417 (2014)

    Article  MathSciNet  Google Scholar 

  33. Queffélec, M.: Substitution Dynamical Systems—Spectral Analysis. 2nd edn. Lecture Notes in Mathematics, vol. 1294. Springer, Berlin (2010)

    Chapter  Google Scholar 

  34. Rao, H., Wen, Z.-Y., Yang, Y.-M.: Dual systems of algebraic iterated function systems. Adv. Math. 253, 63–85 (2014)

    Article  MathSciNet  Google Scholar 

  35. Rauzy, G.: Nombres algébriques et substitutions. Bull. Soc. Math. Fr. 110(2), 147–178 (1982)

    Article  Google Scholar 

  36. Rauzy, G.: Ensembles à restes bornés. In: Seminar on Number Theory, 1983–1984 (Talence, 1983/1984), Exp. No. 24. Université de Bordeaux I, Talence (1984)

  37. Reveillès, J.-P.: Géométrie discrète, calcul en nombres entiers et algorithmique. Thèse de Doctorat, Université Louis Pasteur, Strasbourg (1991)

  38. Sano, Y., Arnoux, P., Ito, S.: Higher dimensional extensions of substitutions and their dual maps. J. Anal. Math. 83, 183–206 (2001)

    Article  MathSciNet  Google Scholar 

  39. Siegel, A., Thuswaldner, J.M.: Topological Properties of Rauzy Fractals. Mémoires de la Société Mathématiques de France, vol. 118. Société Mathématiques de France, Paris (2009)

    Article  Google Scholar 

  40. Solomyak, B.: Pseudo-self-affine tilings in \({\mathbb{R}}^d\). J. Math. Sci. (N.Y.) 140(3), 452–460 (2007)

    Article  MathSciNet  Google Scholar 

  41. Thuswaldner, J.M.: Unimodular Pisot substitutions and their associated tiles. J. Théor. Nombres Bordeaux 18(2), 487–536 (2006)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This research is supported by the ANR/FWF Project “FAN – Fractals and Numeration” (ANR-12-IS01-0002, FWF Grant I1136) of the French National Agency for Research (ANR) and the Austrian Science Fund (FWF) and by the JSPS/FWF Project I3346 “Topology of planar and higher dimensional self-replicating tiles” of the Japan Society for the Promotion of Science (JSPS) and the FWF. The authors are grateful to Jörg Thuswaldner for the precious help and many inspiring discussions which contributed to the creation of this paper, to Valérie Berthé for reading carefully a preliminary version, and to Pierre Arnoux for giving remarkable comments. We also thank warmly the referees: their suggestions increased considerably the quality and readability of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Milton Minervino.

Additional information

Editor in Charge: Kenneth Clarkson

Appendices

Appendix A: Proof of Lemma 5.9

Lemma A.1

Let \(n=p+m\), \(M\in {\mathbb {R}}^{n\times n}\) and \(\{\mathbf{e}_i\}_{i=1,\ldots ,n}\) be the canonical basis of \({\mathbb {R}}^n\). Moreover, let \(\underline{a}=a_1\wedge \cdots \wedge a_m\in {\mathcal {O}}_m\) and \(\mathbf{x}_1,\ldots ,\mathbf{x}_p\in {\mathbb {R}}^n\). Then

$$\begin{aligned}&\det (M\mathbf{x}_1,\ldots ,M\mathbf{x}_p,\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}) \\&\quad =\sum _{\underline{b}=b_1\wedge \cdots \wedge b_{m}\in {\mathcal {O}}_m} \Big (\bigwedge _{i=1}^{p}M_\sigma \Big )_{\underline{a}^*,\underline{b}^*}(-1)^{\underline{a}+\underline{b}} \det (\mathbf{x}_1,\ldots ,\mathbf{x}_{p},\mathbf{e}_{b_1},\ldots ,\mathbf{e}_{b_{m}}). \end{aligned}$$

Proof

We write \(\mathbf{x}_k=\sum _{j=1}^n\mathbf{x}_k^{(j)}{} \mathbf{e}_j\) for \(k\in \{1,\ldots ,p\}\), \(M=(m_{ij})_{1\le i,j\le n}\) and compute the determinant:

$$\begin{aligned}&\det (M\mathbf{x}_1,\ldots ,M\mathbf{x}_p,\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m})\\&= \det \Big (\sum _{i_1=1}^n\sum _{j_1=1}^nm_{i_1j_1}\mathbf{x}_1^{(j_1)}{} \mathbf{e}_{i_1},\ldots , \sum _{i_p=1}^n\sum _{j_p=1}^nm_{i_pj_p}{} \mathbf{x}_p^{(j_p)}{} \mathbf{e}_{i_p}, \mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}\Big ) \\&=\sum _{1\le j_1,\ldots ,j_p\le n}{} \mathbf{x}_1^{(j_1)}\cdots \mathbf{x}_p^{(j_p)} \det \Big (\sum _{i_1=1}^nm_{i_1j_1}\mathbf{e}_{i_1},\ldots ,\sum _{i_p=1}^nm_{i_pj_p}{} \mathbf{e}_{i_p},\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}\Big )\\&=\sum _{\begin{array}{c}1\le j_1,\ldots ,j_p\le n,\\ \#\,\{j_1,\ldots , j_p\}=p\end{array}}{} \mathbf{x}_1^{(j_1)}\cdots \mathbf{x}_p^{(j_p)} \det \Big (\sum _{i_1=1}^nm_{i_1j_1}\mathbf{e}_{i_1},\ldots ,\sum _{i_p=1}^nm_{i_pj_p}{} \mathbf{e}_{i_p},\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}\Big ). \end{aligned}$$

Remember that \(\big (\bigwedge _{i=1}^{p}M\big )_{\underline{a}^*,\underline{b}^*}\) is the \(p\times p\) minor of M obtained by deleting the rows \(a_1,\ldots ,a_m\) and the columns \(b_1,\ldots ,b_m\) of M, where \(\underline{a}=a_1\wedge \cdots \wedge a_m\) and \(\underline{b}=b_1\wedge \cdots \wedge b_m\). Therefore, by definition,

$$\begin{aligned} \det \Big (\sum \limits _{i_1=1}^nm_{i_1j_1}\mathbf{e}_{i_1},\ldots ,\sum _{i_p=1}^nm_{i_pj_p}{} \mathbf{e}_{i_p},\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}\Big )\\ =(-1)^{\underline{a}+((p+1)+\cdots +n)} \Big (\bigwedge _{i=1}^{p}M\Big )_{\underline{a}^*,\underline{j}'^*} {\mathrm{sgn}}(\tau ), \end{aligned}$$

where \(\underline{j}'=j_1'\wedge \cdots \wedge j_m'\in {\mathcal {O}}_m\) such that \(\{j_1,\ldots ,j_p\}\cup \{j_1',\ldots ,j_m'\}=\{1,\ldots ,n\}\), \(\tau \) is the permutation of \(\{1,\ldots ,n\}\) satisfying \(\tau (j_1')=j_1',\ldots ,\tau (j_m')=j_m'\) and \(\tau (j_1)\le \cdots \le \tau (j_p)\) and \({\mathrm{sgn}}(\tau )\) its signature. We denote the set of permutations on \(\{1,\ldots ,n\}\) by \({\mathrm{Per}}_n\). In this way, we obtain

$$\begin{aligned}&\det (M\mathbf{x}_1,\ldots ,M\mathbf{x}_p,\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}) \\&\quad = \sum _{\begin{array}{c}1\le j_1,\ldots ,j_p\le n,\\ \#\,\{j_1,\ldots ,j_p\}=p\end{array}}\Big (\bigwedge _{i=1}^{p}M\Big )_{\underline{a}^*,\underline{j}'^*}(-1)^{\underline{a}+((p+1)+\cdots +n)} {\mathrm{sgn}}(\tau )\mathbf{x}_1^{(j_1)}\cdots \mathbf{x}_p^{(j_p)}\\&\quad = \sum \limits _{\underline{j}=j_1\wedge \ldots \wedge j_p\in {\mathcal {O}}_p}\Big (\bigwedge _{i=1}^{p}M\Big )_{\underline{a}^*,\underline{j}'^*}(-1)^{\underline{a}+((p+1)+\cdots +n)}\\&\qquad \times \sum _{\begin{array}{c}\tau \in {\mathrm{Per}}_n,\\ \tau (j_k')=j_k'\\ (1\le k\le m)\end{array}}{\mathrm{sgn}}(\tau )\mathbf{x}_1^{(\tau (j_1))}\cdots \mathbf{x}_p^{(\tau (j_p))}. \end{aligned}$$

Note that

$$\begin{aligned}&\sum _{\begin{array}{c}\tau \in {\mathrm{Per}}_n,\\ \tau (j_k')=j_k'\\ (1\le k\le m)\end{array}}{\mathrm{sgn}}(\tau )\mathbf{x}_1^{(\tau (j_1))}\cdots \mathbf{x}_p^{(\tau (j_p))}\\&\quad =(-1)^{\underline{j}'+((p+1)+\cdots +n)}\det (\mathbf{x}_1,\ldots ,\mathbf{x}_p,\mathbf{e}_{j_1'},\ldots ,\mathbf{e}_{j_m'}), \end{aligned}$$

which leads to the desired equality, after renaming \(\underline{j}'\leftrightarrow \underline{b}\):

$$\begin{aligned}&\det (M\mathbf{x}_1,\ldots ,M\mathbf{x}_p,\mathbf{e}_{a_1},\ldots \mathbf{e}_{a_m}) \\&\quad =\sum \limits _{\underline{b}=b_1\wedge \ldots \wedge b_m\in {\mathcal {O}}_m}\Big (\bigwedge _{i=1}^{p}M\Big )_{\underline{a}^*,\underline{b}^*}(-1)^{\underline{a}+\underline{b}} \det (\mathbf{x}_1,\ldots ,\mathbf{x}_p,\mathbf{e}_{b_1},\ldots ,\mathbf{e}_{b_m}).\qquad \qquad \end{aligned}$$

\(\square \)

Proof of Lemma 5.9

Let \(\mathbf{u}_{d+1},\ldots ,{\mathbf {u}}_{n}\) be a basis of \({\mathbb {K}}_n\) made of eigenvectors of \(M_\sigma \) for the associated eigenvalues \(\zeta _{d+1},\ldots ,\zeta _n\). Also, let \(\mathbf{v}_{d+1},\ldots ,{\mathbf {v}}_{n}\) be eigenvectors of \({}^tM_\sigma \) for the associated eigenvalues \(\zeta _{d+1},\ldots ,\zeta _n\), normalized to have \({\mathbf {u}}_k\cdot {\mathbf {v}}_k=1\) for \(k\in \{d+1,\ldots ,n\}\). Since \({\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_n\) are all orthogonal to the contracting space \({\mathbb {K}}_c\), the following equality holds for every \(\underline{a}=a_1\wedge \cdots \wedge a_{d-1}\in {\mathcal {O}}_{d-1}\):

$$\begin{aligned} \lambda _{d-1}(\pi _c\overline{({\mathbf {0}},\underline{a})})= \frac{\left| \det \big ({\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_{n},\pi _c(\mathbf{e}_{a_1}),\ldots ,\pi _c(\mathbf{e}_{a_{d-1}})\big )\right| }{\lambda _{\bar{n}}\big ({\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_{n}\big )}, \end{aligned}$$
(A.1)

where \(\lambda _p\) denotes the Lebesgue measure on the p-dimensional space and

$$\begin{aligned} \lambda _{\bar{n}}\left( {\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_{n}\right) =:1/K_0 \end{aligned}$$

is the measure of the parallelotope generated by the vectors \({\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_{n}\).

Note that for all \(\mathbf{x}\in {\mathbb {R}}^n\), we have the decomposition

$$\begin{aligned} \mathbf{x}=\langle \mathbf{x},\mathbf{v}_1\rangle \mathbf{u}_1+\sum _{i=d+1}^n\langle \mathbf{x},\mathbf{v}_i\rangle \mathbf{u}_i+\pi _c(\mathbf{x}). \end{aligned}$$
(A.2)

We use the above decomposition for the vectors \({\mathbf {v}}_1,{\mathbf {v}}_{d+1},\ldots ,{\mathbf {v}}_{\bar{n}}\) and expand the determinant by multilinearity to obtain:

$$\begin{aligned} \begin{array}{rcl} \lambda _{d-1}(\pi _c\overline{({\mathbf {0}},\underline{a})})&{}=&{} \underbrace{K_0\det \big (\langle \mathbf{v}_k,\mathbf{v}_l\rangle _{k,l=1,d+1,\ldots ,n}\big )}_{=:K}\\ &{}&{}\cdot \;\big |\det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\pi _c(\mathbf{e}_{a_1}),\ldots ,\pi _c(\mathbf{e}_{a_{d-1}}))\big |. \end{array} \end{aligned}$$

The terms containing \(\pi _c(\mathbf{v}_k)\) vanished for each value of \(k\in \{1,d+1,\ldots ,n\}\)), since the d vectors \(\pi _c(\mathbf{v}_k),\pi _c(\mathbf{e}_{a_1}),\ldots ,\pi _c(\mathbf{e}_{a_{d-1}})\) are linearly dependent in the \((d-1)\)-dimensional space \({\mathbb {K}}_c\). Using now (A.2) for the vectors \(\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}}\), we can simplify the last expression to

$$\begin{aligned} \lambda _{d-1}(\pi _c\overline{({\mathbf {0}},\underline{a})})= K\big |\det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}})\big |. \end{aligned}$$
(A.3)

Now, by (N), \(g(0)=1=\Pi _{k=d+1}^n\zeta _k\) and therefore

$$\begin{aligned}&\det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}}) \\&\quad = \frac{1}{\beta }\det (\beta \mathbf{u}_1,\zeta _{d+1}\mathbf{u}_{d+1},\ldots ,\zeta _n\mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}})\\ \\&\quad =\frac{1}{\beta }\det (M_\sigma \mathbf{u}_1,M_\sigma \mathbf{u}_{d+1},\ldots ,M_\sigma \mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}}). \end{aligned}$$

By Lemma A.1, we can relate the last determinant to \(\bigwedge _{i=1}^{\bar{n}}M_\sigma \): for all \(\underline{a}=a_1\wedge \cdots \wedge a_{d-1}\in {\mathcal {O}}_{d-1}\), we have

$$\begin{aligned}&\det (M_\sigma \mathbf{u}_1,M_\sigma \mathbf{u}_{d+1},\ldots ,M_\sigma \mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}}) \\&\quad =\sum \limits _{\underline{b}=b_1\wedge \cdots \wedge b_{d-1}\in {\mathcal {O}}_{d-1}} \Big (\bigwedge _{i=1}^{\bar{n}}M_\sigma \Big )_{\underline{a}^*,\underline{b}^*}(-1)^{\underline{a}+\underline{b}} \det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{b_1},\ldots ,\mathbf{e}_{b_{d-1}})\\&\quad =\sum \limits _{\underline{b}\in {\mathcal {O}}_{d-1}} \Big (\bigwedge _{i=1}^{\bar{n}}{}^tM_\sigma \Big )_{\underline{b}^*,\underline{a}^*}(-1)^{\underline{a}^*+\underline{b}^*} \det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{b_1},\ldots ,\mathbf{e}_{b_{d-1}})\\&\quad =\sum \limits _{\underline{b}\in {\mathcal {O}}_{d-1}}(M_{d-1})_{\underline{b},\underline{a}}\det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{b_1},\ldots ,\mathbf{e}_{b_{d-1}}). \end{aligned}$$

We used here that \((-1)^{\underline{a}+\underline{b}}=(-1)^{\underline{a}^*+\underline{b}^*}\). By the above computation, this means that the vector

$$\begin{aligned} \mathbf{V}=\big (\det (\mathbf{u}_1,\mathbf{u}_{d+1},\ldots ,\mathbf{u}_{n},\mathbf{e}_{a_1},\ldots ,\mathbf{e}_{a_{d-1}})\big )_{\underline{a}\in {\mathcal {O}}_{d-1}} \end{aligned}$$

is an eigenvector of \({}^tM_{d-1}\) for the eigenvalue \(\beta \). It follows that

$$\begin{aligned} \beta |{\mathbf {V}}|=|{}^tM_{d-1}{\mathbf {V}}|\le |{}^tM_{d-1}||{\mathbf {V}}|. \end{aligned}$$

Now, it follows from (4.5) and (P) that \(|{}^tM_{d-1}|=\bigwedge _{i=1}^{\bar{n}}M_\sigma \) is a primitive matrix. Therefore, by Lemma 5.8, \(\beta \) is its Perron–Frobenius eigenvalue. Consequently, the inequality \(\beta |{\mathbf {V}}|\le |{}^tM_{d-1}||{\mathbf {V}}|\) is an equality and

$$\begin{aligned} \big (\lambda _{d-1}(\pi _c\overline{({\mathbf {0}},\underline{a})})\big )_{\underline{a}\in {\mathcal {O}}_{d-1}}= K|\mathbf{V}| \end{aligned}$$

is an eigenvector of \(|{}^tM_{d-1}|\) for the eigenvalue \(\beta \). \(\square \)

Appendix B: Proof of Lemma 7.1

The classical Rauzy fractal subtiles \(\mathcal {R}(a)\) can be described via Dumont–Thomas numeration (see [17]) as

$$\begin{aligned} \mathcal {R}(a) = \Bigg \{\sum _{i\ge 0} \pi _c (M_\sigma ^i\,{\mathbf {l}}(p_i)): (p_i)_{i\ge 0} \in {\mathcal {G}}_p(a)\Bigg \}, \end{aligned}$$
(B.1)

where \({\mathcal {G}}_p(a)\) denotes the set of labels of infinite paths in the prefix graph of the substitution ending at \(a\in {\mathcal {A}}\) (for more details see e.g. [13, 15]). We can follow as well infinite paths in the suffix graph. If we do this we get instead

$$\begin{aligned} -\mathcal {R}(a)-\pi _c({\mathbf {e}}_a) = \Bigg \{\sum _{i\ge 0} \pi _c (M_\sigma ^i\,{\mathbf {l}}(s_i)): (s_i)_{i\ge 0} \in {\mathcal {G}}_s(a)\Bigg \}, \end{aligned}$$
(B.2)

where \({\mathcal {G}}_s(a)\) denotes the set of labels of infinite paths in the suffix graph of the substitution ending at \(a\in {\mathcal {A}}\) (see [15, Sect. 5]).

We can use Dumont–Thomas numeration to describe also the subtiles \({\mathcal {R}}(a\wedge b)\) using the \({\mathbf {E}}^2(\sigma )\) -suffix graph that is defined as follows.

Definition B.1

The \({\mathbf {E}}^2(\sigma )\)-suffix graph has set of vertices \(\{\underline{a}^*:\underline{a}\in \wedge ^3{\mathcal {A}}\}\), and there is an edge \(\underline{a}^* \xrightarrow {\mathbf {s}}\underline{b^*}\) if and only if \(\sigma (\underline{a}) = {\mathbf {p}}\underline{b}\mathbf {s}\), or equivalently if and only if \((M_\sigma ^{-1}{\mathbf {l}}(\mathbf {s}),\underline{a}^*) \in {\mathbf {E}}^2(\sigma )({\mathbf {0}},\underline{b}^*)\).

In Fig. 15 the \({\mathbf {E}}^2(\sigma _t)\)-suffix graph is depicted. Observe that this graph with reversed edges describes the images of every face by \({\mathbf {E}}^2(\sigma _t)\).

Fig. 15
figure 15

The \({\mathbf {E}}^2(\sigma _t)\)-suffix graph. Note that an edge labeled by \(5,\ldots ,1^{t-1}5\) denotes that there exist t edges, one labeled by 5, one by 15, and so on, until \(1^{t-1}5\) (if \(t=0\) then there is no edge of this type). The same is valid for the edges labeled by \(2,\ldots ,1^{t}2\) (for \(t=0\) there is only one edge labeled by 2)

Proposition B.2

We have

$$\begin{aligned} \mathcal {R}(a\wedge b) = \Bigg \{ \sum _{i\ge 0}\pi _c(M_\sigma ^i {\mathbf {l}}(\mathbf {s}_i)) : (\mathbf {s}_i)_{i\ge 0} \in {\mathcal {G}}_{s}(a\wedge b)\Bigg \} \end{aligned}$$

where \({\mathcal {G}}_s(a\wedge b)\) denotes the set of labels of infinite walks in the \({\mathbf {E}}^2(\sigma )\)-suffix graph ending at state \(a\wedge b\).

Proof

This is a direct consequence of Proposition 6.2 and of the definition of \({\mathbf {E}}^2(\sigma )\). \(\square \)

By abuse of notation we will write 0 instead of \(\epsilon \) by reading labels of walks in the suffix or \({\mathbf {E}}^2(\sigma _0)\)-suffix graphs.

We will relate the elements \(\sum _{i\ge 0} \pi _c(M_\sigma ^i {\mathbf {l}}(\mathbf {s}_i))\) with \((\mathbf {s}_i)_{i\ge 0}\in {\mathcal {G}}_{s}(a\wedge b)\) with those \(\sum _{i\ge 0} \pi _c(M_\sigma ^i {\mathbf {l}}(\mathbf {s}_i))\) for \((\mathbf {s}_i)_{i\ge 0} \in {\mathcal {G}}_s(a)\).

For the Hokkaido substitution \(\sigma _0\) we have

$$\begin{aligned} {\mathcal {G}}_s(a) = \big \{ (\mathbf {s}_i)_{i\ge 0} = \cdots 0^5 2^{k_2} 0^5 2^{k_1} 0^{a-1} : 0\le k_i \le \infty \big \}. \end{aligned}$$

Proof of Lemma  7.1

Observe that

$$\begin{aligned} \pi _e(M_\sigma ^3{\mathbf {e}}_2) = \pi _e(M_\sigma {\mathbf {e}}_2) + \pi _e({\mathbf {e}}_2) \quad \text {i.e.}\quad \delta (2000.) = \delta (0022.), \end{aligned}$$
(B.3)

where \(\delta (w) = \sum _{i=0}^{|w|} \pi _e(M_\sigma ^i {\mathbf {l}}(w_i))\). Notice that we can extend \(\delta \) to infinite strings \((s_i)_{i\ge 0}\). We will prove using (B.3) that \(\delta (\mathcal {G}_s(2\wedge 3))=\delta ({\mathcal {G}}_s(1)\cup {\mathcal {G}}_s(4))\). For this reason we will write \(w =_\delta w'\) if \(\delta (w)=\delta (w')\). The cycle

figure b

in the graph of Fig. 15 produces strings of type \(0020^3(20)^k002=_\delta 0^5 2^{2k+2}\). Starting from state \(2\wedge 5\) we get strings \(0^5 2^{2k+1}\). Walking from the first node \(2\wedge 5\) to the second \(2\wedge 5\) returns \(0^5 2^{2k}\) and extending this walk to the left starting from \(2\wedge 3\) we obtain \(0^5 2^{2k+3}\). Walking in Fig. 15 from \(2\wedge 3\) to \(2\wedge 5\) we get the word \(0022002=_\delta 2 0^5 2\). Thus, \(\delta (\mathcal {G}_s(2\wedge 3)) = \delta (\cdots 0^52^{k_2}0^52^{k_1})\), with \(0\le k_i\le \infty \), i.e., \(\delta ({\mathcal {G}}_s(1))\subseteq \delta (\mathcal {G}_s(2\wedge 3))\). Strings ending with \(0^3\) are obtained following the loop \(2\wedge 3 {\mathop {\rightarrow }\limits ^{00}}4\wedge 5{\mathop {\rightarrow }\limits ^{2}}2\wedge 5{\mathop {\rightarrow }\limits ^{2}}2\wedge 3\). Since these are all possible non-trivial paths ending at \(2\wedge 3\) we have proven that \(\delta (\mathcal {G}_s(2\wedge 3))=\delta ({\mathcal {G}}_s(1)\cup {\mathcal {G}}_s(4))\) which implies \({\mathcal {R}}(2\wedge 3) = (-{\mathcal {R}}(1)-\pi _c({\mathbf {e}}_1)) \cup (-{\mathcal {R}}(4)-\pi _c({\mathbf {e}}_4))\) by (B.2).

Since \(2\wedge 3\) goes to \(3\wedge 4\) by reading a 0 we deduce immediately that all the strings ending at \(3\wedge 4\) are equivalent under \(\delta \) to those in \({\mathcal {G}}_s(2)\cup {\mathcal {G}}_s(5)\). Hence by (B.2) we get \({\mathcal {R}}(3\wedge 4) = (-{\mathcal {R}}(2)-\pi _c({\mathbf {e}}_2)) \cup (-{\mathcal {R}}(5)-\pi _c({\mathbf {e}}_5))\).

Starting from \(2\wedge 5\) and going to \(2\wedge 4\) passing by \(1\wedge 3\) we read 00 and by the above reasonings we get then all possible strings \(\cdots 2^{k_2}0^52^{k_1}00\) belonging to \({\mathcal {G}}_s(3)\). From \(2\wedge 5 {\mathop {\rightarrow }\limits ^{00}} 2\wedge 4{\mathop {\rightarrow }\limits ^{0}}3\wedge 5{\mathop {\rightarrow }\limits ^{2}}2\wedge 4\) we get all expansions in \({\mathcal {G}}_s(5)+2\), where with the latter we mean the set of \((s_i)_{i\ge 0}\in {\mathcal {G}}_s(5)\) such that \(s_0=2\). Walking k times through the loop \(2\wedge 4\rightarrow 3\wedge 5 \rightarrow 2\wedge 4\) and extending to the left with \(2\wedge 5\) we get strings \(0^2 (02)^k\). Subtracting \(v_4 = \delta (200.)\) we get \(0^2(02)^{k-2}0002 =_\delta 0^5 2^{2k-3}\). Walking through the loop \(2\wedge 4 \rightarrow 3\wedge 5 \rightarrow 2\wedge 5 \rightarrow 2\wedge 4\) and then once into \(2\wedge 4\rightarrow 3\wedge 5 \rightarrow 2\wedge 4\) we read the string \(020^5 =_\delta 0^42^3\) and, after subtracting \(v_4\), we get \(0^5 2^2\). Repeating this loop we get arbitrary large strings ending with an even number of 2s. Thus we have shown we get strings in \({\mathcal {G}}_s(1)+4\).

So by (B.2) we just proved that \({\mathcal {R}}(2\wedge 4)\) is made of the domains \(\delta ({\mathcal {G}}_s(3)) = -{\mathcal {R}}(3)-\pi _c({\mathbf {e}}_3)\), \(\delta ({\mathcal {G}}_s(5)+2) = -{\mathcal {R}}(5)-\pi _c({\mathbf {e}}_5)+\pi _c({\mathbf {e}}_2) = -{\mathcal {R}}(5)-\pi _c({\mathbf {e}}_3)\) and \(\delta ({\mathcal {G}}_s(1)+4) = -{\mathcal {R}}(1)-\pi _c({\mathbf {e}}_1) + \pi _c({\mathbf {e}}_4) = -{\mathcal {R}}(1)-\pi _c({\mathbf {e}}_3)\), since \(\pi _c({\mathbf {e}}_1)=\pi _c({\mathbf {e}}_3)+\pi _c({\mathbf {e}}_4)\) and \(\pi _c({\mathbf {e}}_5)=\pi _c({\mathbf {e}}_2)+\pi _c({\mathbf {e}}_3)\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Loridant, B., Minervino, M. Geometrical Models for a Class of Reducible Pisot Substitutions. Discrete Comput Geom 60, 981–1028 (2018). https://doi.org/10.1007/s00454-018-9969-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00454-018-9969-0

Keywords

Mathematics Subject Classification

Navigation