Skip to main content
Log in

Renormalisation of Pair Correlation Measures for Primitive Inflation Rules and Absence of Absolutely Continuous Diffraction

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

The pair correlations of primitive inflation rules are analysed via their exact renormalisation relations. We introduce the inflation displacement algebra that is generated by the Fourier matrix of the inflation and deduce various consequences of its structure. Moreover, we derive a sufficient criterion for the absence of absolutely continuous diffraction components, as well as a necessary criterion for its presence. This is achieved via estimates for the Lyapunov exponents of the Fourier matrix cocycle of the inflation rule. We also discuss some consequences for the spectral measures of such systems. While we develop the theory first for the classic setting in one dimension, we also present its extension to primitive inflation rules in higher dimensions with finitely many prototiles up to 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

Similar content being viewed by others

Notes

  1. Note that there is another topological dynamical system, denoted by \((\mathbb {X}, {\mathbb {Z}})\), which emerges from the shift action on the symbolic hull \(\mathbb {X}\), the latter obtained as the orbit closure of a symbolic fixed point of \(\varrho \) or a suitable power of it; see [5] for more. Also this system is strictly ergodic.

  2. This notion of irreducibility is to be distinguished from the one for non-negative matrices used elsewhere in this paper. Since this will always be clear from the context, we stick to the standard terminology.

  3. We use \(\varSigma _n\) for the symmetric or permutation group of n symbols.

  4. A square matrix M is called decomposable if it can be brought to the block-triangular form \(M' = \left( {\begin{matrix} A &{} 0 \\ B &{} C \end{matrix}} \right) \) via simultaneous permutations of its rows and columns, and indecomposable otherwise [34].

  5. As we outline in more detail in the Appendix, one can alternatively work with the recursion from Lemma 3.26 directly. However, the dimensional reduction leads to a stronger result in the sense that we also get a representation of \(\varvec{h}\) that resembles the situation of the pure point part.

  6. We refer to Example 5.12 in the next section for an illustration.

  7. Some statistical data, such as the relative frequency of the prototiles or the frequency of vertices that carry a control point, can be extracted with standard Perron–Frobenius arguments from either rule.

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, Birkhäuser, Basel (2015)

  2. Baake, M., Frank, N.P., Grimm, U., Robinson, E.A.: Geometric properties of a binary non-Pisot inflation and absence of absolutely continuous diffraction. Stud. Math. 247, 109–154 (2019). arXiv:1706.03976

  3. Baake, M., Gähler, F.: Pair correlations of aperiodic inflation rules via renormalisation: some interesting examples. Topol. Appl. 205, 4–27 (2016). arXiv:1511.00885

  4. Baake, M., Grimm, U.: Squirals and beyond: substitution tilings with singular continuous spectrum. Ergod. Theory Dyn. Syst. 34, 1077–1102 (2014). arXiv:1205.1384

  5. Baake, M., Grimm, U.: Aperiodic Order, vol. 1: A Mathematical Invitation. Cambridge University Press, Cambridge (2013)

  6. Baake, M., Grimm, U. (eds.): Aperiodic Order, vol. 2: Crystallography and Almost Periodicity. Cambridge University Press, Cambridge (2017)

  7. Baake, M., Grimm, U.: Renormalisation of pair correlations and their Fourier transforms for primitive block substitutions. In: Akiyama, S., Arnoux, P. (eds.), Tiling and Discrete Geometry, Springer, Berlin (in press) (2018). arXiv:1906.10484

  8. Baake, M., Grimm, U., Mañibo, N.: Spectral analysis of a family of binary inflation rules. Lett. Math. Phys. 108, 1783–1805 (2018). arXiv:1709.09083

  9. Baake, M., Haynes, A., Lenz, D.: Averaging almost periodic functions along exponential sequences. In: Baake, M., Grimm, U. (eds.): Aperiodic Order, vol. 2: Crystallography and Almost Periodicity, pp. 343–362. Cambridge University Press, Cambridge (2017)

  10. Baake, M., Lenz, D.: Dynamical systems on translation bounded measures: pure point dynamical and diffraction spectra. Ergod. Theory Dyn. Syst. 24, 1867–1893 (2004). arXiv:math.DS/0302231

  11. Baake, M., Lenz, D.: Spectral notions of aperiodic order. Discr. Cont. Dynam. Syst. S 10, 161–190 (2017). arXiv:1601.06629

  12. Baake, M., Lenz, D., van Enter, A.C.D.: Dynamical versus diffraction spectrum for structures with finite local complexity. Ergod. Theory Dyn. Syst. 35, 2017–2043 (2015). arXiv:1307.7518

  13. Baake, M., Moody, R.V.: Weighted Dirac combs with pure point diffraction. J. Reine Angew. Math. (Crelle) 573, 61–94 (2004). arXiv:math.MG/0203030

  14. Barreira, L., Pesin, Y.: Nonuniform Hyperbolicity. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  15. Bartlett, A.: Spectral theory of \({\mathbb{Z}}^d\) substitutions. Ergod. Theory Dyn. Syst. 38, 1289–1341 (2018). arXiv:1410.8106

  16. Berg, C., Forst, G.: Potential Theory on Locally Compact Abelian Groups. Springer, Berlin (1975)

    Book  MATH  Google Scholar 

  17. Berlinkov, A., Solomyak, B.: Singular substitutions of constant length. Ergod. Theory Dyn. Syst. (in press). arXiv:1705.00899

  18. Bufetov, A.I., Solomyak, B.: On the modulus of continuity for spectral measures in substitution dynamics. Adv. Math. 260, 84–129 (2014). arXiv:1305.7373

  19. Bufetov, A.I., Solomyak, B.: A spectral cocycle for substitution systems and translation flows. (preprint). arXiv:1802.04783

  20. Chan, L., Grimm, U.: Spectrum of a Rudin–Shapiro-like sequence. Adv. Appl. Math. 87, 16–23 (2017). arXiv:1611.04446

  21. Chan, L., Grimm, U., Short, I.: Substitution-based structures with absolutely continuous spectrum. Indag. Math. 29, 1072–1086 (2018). arXiv:1706.05289

  22. Clark, A., Sadun, L.: When size matters: subshifts and their related tiling spaces. Ergod. Theory Dyn. Syst. 23, 1043–57 (2003). arXiv:math.DS/0201152

  23. Damanik, D., Sims, R., Stolz, G.: Localization for one-dimensional, continuum, Bernoulli–Anderson models. Duke Math. J. 114, 59–100 (2002). arXiv:math-ph/0010016

  24. Dekking, F.M.: The spectrum of dynamical systems arising from substitutions of constant length. Z. Wahrscheinlichkeitsth. Verw. Geb. 41, 221–239 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  25. Durand, F.: A characterization of substitutive sequences using return words. Discr. Math. 179, 89–101 (1998). arXiv:0807.3322

  26. Einsiedler, M., Ward, T.: Ergodic Theory—With a View Towards Number Theory, GTM 259. Springer, London (2011)

    MATH  Google Scholar 

  27. Everest, G., Ward, T.: Heights of Polynomials and Entropy in Algebraic Dynamics. Springer, London (1999)

    Book  MATH  Google Scholar 

  28. Fan, A.-H., Saussol, B., Schmeling, J.: Products of non-stationary random matrices and multiperiodic equations of several scaling factors. Pac. J. Math. 214, 31–54 (2004). arXiv:math.DS/0210347

  29. Frank, N.P.: Substitution sequences in \({\mathbb{Z}}^d\) with a nonsimple Lebesgue component in the spectrum. Ergod. Theory Dyn. Syst. 23, 519–532 (2003)

    Article  Google Scholar 

  30. Frank, N.P.: Multi-dimensional constant-length substitution sequences. Topol. Appl. 152, 44–69 (2005)

    Article  MATH  Google Scholar 

  31. Frank, N.P.: Introduction to hierarchical tiling dynamical systems. In: Akiyama, S, Arnoux, P. (eds.), Tiling and Discrete Geometry, Springer, Berlin (in press; preprint) (2018). arXiv:1802.09956

  32. Frettlöh, D.: More inflation tilings. In: Baake, M., Grimm, U. (eds.): Aperiodic Order, vol. 2: Crystallography and Almost Periodicity, pp. 1–37. Cambridge University Press, Cambridge (2017)

  33. Frettlöh, D., Richard, C.: Dynamical properties of almost repetitive Delone sets. Discr. Contin. Dyn. Syst. A 34, 531–556 (2014). arXiv:1210.2955

  34. Gantmacher, F.: Matrizentheorie. Springer, Berlin (1986)

    Book  Google Scholar 

  35. Gil de Lamadrid, J., Argabright, L.N.: Almost Periodic Measures, Memoirs AMS, vol. 428. AMS, Providence (1990)

  36. Godrèche, C., Lançon, F.: A simple example of a non-Pisot tiling with fivefold symmetry. J. Phys. I (Fr.) 2, 207–220 (1992)

    Article  Google Scholar 

  37. Hof, A.: On diffraction by aperiodic structures. Commun. Math. Phys. 169, 25–43 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. James, G., Liebeck, M.: Representations and Characters of Groups, 2nd edn. Cambridge University Press, Cambridge (2001)

    Book  MATH  Google Scholar 

  39. Kellendonk, J., Lenz, D., Savinien, J. (eds.): Mathematics of Aperiodic Order. Birkhäuser, Basel (2015)

    MATH  Google Scholar 

  40. Lançon, F., Billiard, L.: Two-dimensional system with a quasicrystalline ground state. J. Phys. (Fr.) 49, 249–256 (1988)

    Article  MathSciNet  Google Scholar 

  41. Lang, S.: Algebra, 3rd, revised edn. Springer, New York (2002)

    Book  MATH  Google Scholar 

  42. Lee, J.-Y., Moody, R.V., Solomyak, B.: Pure point dynamical and diffraction spectra. Ann. Henri Poincaré 3, 1003–1018 (2002). arXiv:0910.4809

  43. Lenz, D.: Continuity of eigenfunctions of uniquely ergodic dynamical systems and intensity of Bragg peaks. Commun. Math. Phys. 287, 225–258 (2009). arXiv:math-ph/0608026

  44. Lenz, D., Strungaru, N.: Pure point spectrum for measure dynamical systems on locally compact Abelian groups. J. Math. Pures Appl. 92, 323–341 (2009). arXiv:0704.2498

  45. Lomonosov, V., Rosenthal, P.: The simplest proof of Burnside’s theorem on matrix algebras. Linear Algebra Appl. 383, 45–47 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  46. Mañibo, N.: Lyapunov exponents for binary substitutions of constant length. J. Math. Phys. 58(113504), 1–9 (2017). arXiv:1706.00451

  47. Mañibo, N.: Spectral analysis of primitive inflation rules. Oberwolfach Rep. 14, 2830–2832 (2017)

    Google Scholar 

  48. Mañibo, N.: Lyapunov Exponents in the Spectral Theory of Primitive Inflation Systems. PhD Thesis, Bielefeld University (2019). https://pub.uni-bielefeld.de/record/2935972

  49. Moody, R.V., Strungaru, N.: Almost periodic measures and their Fourier transforms. In: Baake, M., Grimm, U. (eds.): Aperiodic Order, vol. 2: Crystallography and Almost Periodicity, pp.173–270. Cambridge University Press, Cambridge (2017)

  50. Müller, P., Richard, C.: Ergodic properties of randomly coloured point sets. Can. J. Math. 65, 349–402 (2013). arXiv:1005.4884

  51. Queffélec, M.: Substitution Dynamical Systems: Spectral Analysis, 2nd edn., LNM 1294, Springer, Berlin (2010)

  52. Robinson Jr., E.A.: Symbolic dynamics and tilings of \({\mathbb{R}}^{d}\). Proc. Symp. Appl. Math. 60, 81–119 (2004)

    Article  MATH  Google Scholar 

  53. Rudin, W.: Fourier Analysis on Groups. Wiley, New York (1962)

    MATH  Google Scholar 

  54. Scott, W.R.: Group Theory. Prentice Hall, Englewood Cliffs (1964)

    MATH  Google Scholar 

  55. Sing, B.: Pisot Substitutions and Beyond. PhD Thesis, Bielefeld University (2006). https://pub.uni-bielefeld.de/record/2302336

  56. Solomyak, B.: Dynamics of self-similar tilings. Ergod. Theory Dyn. Syst. 17, 695–738 (1997) and Ergod. Theory Dyn. Syst. 19, 1685 (1999) (erratum)

  57. Solomyak, B.: Nonperiodicity implies unique composition for self-similar translationally finite tilings. Discr. Comput. Geom. 20, 265–278 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  58. Strungaru, N.: On the Fourier analysis of measures with Meyer set support (preprint). arXiv:1807.03815

  59. Trebin, H.R. (ed.): Quasicrystals: Structure and Physical Properties. Wiley-VCH, Weinheim (2003)

    Google Scholar 

  60. Viana, M.: Lectures on Lyapunov Exponents. Cambridge University Press, Cambridge (2013)

    MATH  Google Scholar 

  61. Vince, A.: Digit tiling of Euclidean space. In: Baake, M., Moody, R.V. (eds.) Directions in Mathematical Quasicrystals, CRM Monograph Series, vol. 13, pp. 329–370. AMS, Providence (2000)

    Chapter  Google Scholar 

Download references

Acknowledgements

It is a pleasure to thank Frederic Alberti, Alan Bartlett, Scott Balchin, Natalie Frank, Uwe Grimm, Andrew Hubery, Robbie Robinson, Boris Solomyak and Nicolae Strungaru for helpful discussions. We also thank two anonymous reviewers for their thoughtful comments. This work was supported by the German Research Foundation (DFG), within the CRC 1283.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Baake.

Additional information

Communicated by J. Marklof

Publisher's Note

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

Appendix: The (Skew) Kronecker Product Algebra

Appendix: The (Skew) Kronecker Product Algebra

The structure of the correlation measures relies on some properties of the Kronecker product matrices

$$\begin{aligned} \varvec{A} (k) \, {:}{=}\, B(k) \otimes \overline{B(k)} , \end{aligned}$$
(36)

defined for \(k\in \mathbb {R}\). Obviously, one has \(\overline{\varvec{A} (k)} = \varvec{A} (-k)\) and \(\det (\varvec{A} (k)) = |\det (B(k))|^{2 n_{a}}\).

In view of the structure of Eq. (36), let us now consider the \(\mathbb {R}\)-algebra \(\varvec{\mathcal {A}}\) that is generated by the matrix family \(\{ \varvec{A} (k) \mid k\in \mathbb {R}\}\). Due to the Kronecker product structure, \(\varvec{\mathcal {A}}\) fails to be irreducible, no matter what the structure of the IDA \(\mathcal {B}\) is. Let us explore this in some more detail. Let \(V=\mathbb {C}^{n_{a}}\) and consider \(W{:}{=}V \otimes ^{}_{\mathbb {C}} V\), the (complex) tensor product, which is a vector space over \(\mathbb {C}\) of dimension \(n_{a}^{2}\), but also one over \(\mathbb {R}\), then of dimension \(2n_{a}^{2}\). In the latter setting, consider the involution \(C \! : \, W \!\longrightarrow W\) defined by

$$\begin{aligned} x \otimes y \; \longmapsto \; C (x \otimes y) \, {:}{=}\, \overline{y \otimes x} \, = \, {\bar{y}} \otimes {\bar{x}} \end{aligned}$$

together with its unique extension to an \(\mathbb {R}\)-linear mapping on W. Observe that there is no \(\mathbb {C}\)-linear extension, because \(C \bigl (a (x \otimes y)\bigr ) = {\bar{a}}\; C (x\otimes y)\) for \(a \in \mathbb {C}\). With this definition of C, one finds for an arbitrary \(k\in \mathbb {R}\) that

$$\begin{aligned} \begin{aligned} \varvec{A} (k)\, C(x\otimes y) \,&= \, \bigl (B(k) \otimes \,\overline{\! B(k)\! }\, \bigr ) ( {\bar{y}} \otimes {\bar{x}} ) \, = \, \bigl (B(k){\bar{y}}\bigr ) \otimes \bigl (\, \overline{\! B(k)x } \, \bigr ) \\&= \, C \bigl ( \bigl (B(k) \otimes \, \overline{\! B(k)\! }\, \bigr ) (x \otimes y ) \bigr ) \, = \, C \bigl ( \varvec{A} (k) \, ( x \otimes y ) \bigr ), \end{aligned} \end{aligned}$$

so C commutes with the linear map defined by \(\varvec{A} (k)\), for any \(k\in \mathbb {R}\). The \(\mathbb {R}\)-linear mapping C has eigenvalues \(\pm 1\) and is diagonalisable, as follows from the unique splitting of an arbitrary \(w\in W\) as \(w = \frac{1}{2} \bigl ( w + C(w)\bigr ) + \frac{1}{2} \bigl ( w - C(w)\bigr )\). So, our vector space splits as \(W \! = W_{\! +} \oplus W_{\! -}\) into real vector spaces that are eigenspaces of C. Their dimensions are

$$\begin{aligned} \dim ^{}_{\mathbb {R}} (W_{\! +}) \, = \, \dim ^{}_{\mathbb {R}} (W_{\! -}) \, = \, n_{a}^{2} \end{aligned}$$

since \(W_{\! -} = \mathrm {i}W_{\! +}\) with \(W_{\! +} \cap W_{\! -} = \{ 0 \}\). It is thus clear that \(W_{\! +}\) and \(W_{\! -}\) are invariant (real) subspaces for the \(\mathbb {R}\)-algebra \(\varvec{\mathcal {A}}\).

Observe next that we have

$$\begin{aligned} \varvec{A} (k) \, = \, B(k) \otimes \overline{B (k)} \; = \! \sum _{x,y \in S_{T}}\! \mathrm {e}^{2 \pi \mathrm {i}k (x-y)} \, D_{x} \otimes D_{y} \; = \sum _{z\in \triangle _{T}} \mathrm {e}^{2 \pi \mathrm {i}k z}\, F_{z} \end{aligned}$$

where \(\triangle _{T} {:}{=}S_{T} - S_{T}\) is the Minkowski difference, with \(-\triangle _{T}=\triangle _{T}\), and

$$\begin{aligned} F_{z} \; = \sum _{\begin{array}{c} x,y \in S_{T} \\ x-y = z \end{array}} D_{x} \otimes D_{y} . \end{aligned}$$
(37)

In analogy to before, \(F_{z} = F_{z'}\) is possible for \(z\ne z'\). For instance, if \(z=x-y\) with \(x\ne y\) and \(D_{x} = D_{y}\), one can get \(F_{z} = F_{-z}\) if there is no other way to write z as a difference of two numbers in \(S_{T}\).

For \(a\in \mathbb {C}\), one easily checks that \(C \circ (a F_{z}) = {\bar{a}} C \circ F_{z} = {\bar{a}} F_{-z} \circ C\), which implies \([C, \varvec{A}(k)]=0\) for all \(k\in \mathbb {R}\), in line with our previous derivation. It is immediate that \(\varvec{\mathcal {A}}\) is contained in the \(\mathbb {R}\)-algebra \(\varvec{\mathcal {A}}_{F}\) that is generated by the matrices \(\{ F_{z}+F_{-z} \mid 0 \leqslant z\in \triangle _{T}\}\) together with \(\{ \mathrm {i}( F_{z} - F_{-z}) \mid 0 \leqslant z\in \triangle _{T}\}\), and an argument similar to the one previously used for \(\mathcal {B}\) shows that \( \varvec{\mathcal {A}}_{F} \subseteq \varvec{\mathcal {A}}\), hence \(\varvec{\mathcal {A}} = \varvec{\mathcal {A}}_{F}\). Since \(\dim ^{}_{\mathbb {C}} (\mathcal {B}) \leqslant n_{a}^{2}\), and since we generate the real algebra only after taking the Kronecker product, one has

$$\begin{aligned} \dim ^{}_{\mathbb {R}} (\varvec{\mathcal {A}}) \, \leqslant \, n_{a}^{4} , \end{aligned}$$

which is also clear from \(\dim ^{}_{\mathbb {R}} (W_{\! +}) = n_{a}^{2}\). Moreover, one has the following result.

Lemma 5.15

Let \(\varrho \) be a primitive inflation rule on an alphabet with \(n_{a}\) letters, and assume that the IDA \(\mathcal {B}\) of \(\varrho \) is irreducible over \(\mathbb {C}\). Then, the induced \(\mathbb {R}\)-algebra \(\varvec{\mathcal {A}}\) is isomorphic with \({{\,\mathrm{Mat}\,}}(n_{a}^2, \mathbb {R})\), and its action on the subspace \(W_{\! +}\) is irreducible as well, this time over \(\mathbb {R}\).

Proof

Here, \(\mathcal {B}\) irreducible over \(\mathbb {C}\) means \(\mathcal {B}= {{\,\mathrm{Mat}\,}}(n_{a}, \mathbb {C})\). With \(\varGamma {:}{=}\mathcal {B}\otimes ^{}_{\mathbb {C}} \mathcal {B}\), where \(\otimes ^{}_{\mathbb {C}}\) denotes the tensor product over \(\mathbb {C}\), one has \(\varGamma \simeq {{\,\mathrm{Mat}\,}}(n_{a}^{2}, \mathbb {C})\) by standard arguments. Clearly, \(\varGamma \) is a \(\mathbb {C}\)-algebra of dimension \(n_{a}^{4}\), but also an \(\mathbb {R}\)-algebra, then of dimension \(2 n_{a}^{4}\). Now, using the Kronecker product as representation of the tensor product, the mapping defined by \(M\otimes N \mapsto \overline{N}\otimes \overline{M}\) has a unique extension to an automorphism \(\sigma \) of \(\varGamma \) as an \(\mathbb {R}\)-algebra.

Our \(\mathbb {R}\)-algebra \(\varvec{\mathcal {A}}\) consists of all fixed points of \(\sigma \), so \(\varvec{\mathcal {A}} = \{ Q \in \varGamma \mid \sigma (Q) = Q \}\). Employing the elementary matrices \(E_{ij}\) from \({{\,\mathrm{Mat}\,}}(n_{a}, \mathbb {R})\) with \(E_{ij,k\ell } {:}{=}E_{ik} \otimes E_{j \ell }\), we can give a basis of \(\varvec{\mathcal {A}}\), seen as a vector space over \(\mathbb {R}\), by

$$\begin{aligned} \big \{ \tfrac{1}{2} (E_{ij,k\ell } + E_{k\ell ,ij}) \mid (i,j) \leqslant (k,\ell ) \big \} \cup \big \{ \tfrac{\mathrm {i}}{2} (E_{ij,k\ell } - E_{k\ell ,ij}) \mid (i,j) < (k,\ell ) \big \} , \end{aligned}$$

where lexicographic ordering is used for the double indices. Note that the cardinalities are \(\frac{1}{2} n_{a}^{2} (n_{a}^{2} + 1)\) and \(\frac{1}{2} n_{a}^{2} (n_{a}^{2} - 1)\), which add up to \(\dim ^{}_{\mathbb {R}} (\varvec{\mathcal {A}}) = n_{a}^{4}\).

Next, observe that we can get \(E_{ij,k\ell }\) and \(E_{k\ell ,ij}\) by a simple (complex) linear combination of \( \tfrac{1}{2} (E_{ij,k\ell } + E_{k\ell ,ij})\) and \(\tfrac{\mathrm {i}}{2} (E_{ij,k\ell } - E_{k\ell ,ij})\), and vise versa. Put together, this defines a (complex) inner automorphism of \(\varGamma \). Observing that \({{\,\mathrm{Mat}\,}}(n_{a}^{2}, \mathbb {R}) = \{ Q \in \varGamma \mid {\overline{Q}} = Q \}\), this construction can now be used to show that \(\varvec{\mathcal {A}} \simeq {{\,\mathrm{Mat}\,}}(n_{a}^{2}, \mathbb {R})\), which is a central simple algebra. Since \(\varvec{\mathcal {A}} W_{\! +} \subseteq W_{\! +}\) and \(\dim ^{}_{\mathbb {R}} (W_{\! +}) = n_{a}^{2}\), the claimed irreducibility over the reals follows. \(\quad \square \)

Note that all \(F_{z}\) are non-negative, integer matrices. They clearly satisfy the relation \(\sum _{z\in \triangle _{T}} F_{z} = \varvec{A} (0) = M_{\varrho } \otimes M_{\varrho }\). Moreover, under some mild conditions, the spectral radius of \(F^{}_0\) is \(\lambda \), while the other matrices \(F_z\) have smaller spectral radius.

Let us come back to Eq. (36), which implies

$$\begin{aligned} \Vert \varvec{A} (k) \Vert ^{}_{\mathrm {F}} \, = \, \Vert B (k) \Vert ^{2}_{\mathrm {F}} . \end{aligned}$$

If we consider the matrix cocycle defined by \(\varvec{A}^{(n)} (k) = B^{(n)} (k) \otimes \overline{B^{(n)} (k)}\), it is immediate that the maximal Lyapunov exponents, for all \(k\in \mathbb {R}\), are related by

$$\begin{aligned} \chi ^{\varvec{A}} (k) \, = \, 2 \chi ^{B} (k) , \end{aligned}$$
(38)

which also holds for the higher-dimensional case with \(k\in \mathbb {R}^d\). Clearly, one can now reformulate Theorems 3.28 and 5.7 in terms of \(\chi ^{\varvec{A}}\). In particular, one has the following reformulation of Theorem 3.34 and Corollary 3.35 and their higher-dimensional analogues.

Corollary 5.16

Let \(\varrho \) be a primitive inflation rule in \(\mathbb {R}^d\) with finitely many translational prototiles and expansive map Q. Let \(B^{(n)} (.)\) be its Fourier matrix cocycle, with \(\det (B(k))\ne 0\) for at least one \(k\in \mathbb {R}^d\), and \(\varvec{A}^{(n)} (.) = B^{(n)} (.) \otimes \overline{B^{(n)} (.)}\) the corresponding Kronecker product cocycle. If the diffraction measure of the hull defined by \(\varrho \) contains a non-trivial absolutely continuous component, one has \(\chi ^{\varvec{A}} (k) = \log |\det (Q) |\) for a subset of \(\mathbb {R}^d\) of positive measure, which has full measure when \(\chi ^{\varvec{A}} (k)\) is constant for a.e. \(k\in \mathbb {R}^d\). \(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Baake, M., Gähler, F. & Mañibo, N. Renormalisation of Pair Correlation Measures for Primitive Inflation Rules and Absence of Absolutely Continuous Diffraction. Commun. Math. Phys. 370, 591–635 (2019). https://doi.org/10.1007/s00220-019-03500-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03500-w

Navigation