Skip to main content
Log in

An averaging formula for Nielsen numbers on infra-solvmanifolds

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

Abstract

Until now only for special classes of infra-solvmanifolds, namely, infra-nilmanifolds and infra-solvmanifolds of type (R), there was a formula available for computing the Nielsen number of a self-map on those manifolds. In this paper, we provide a general averaging formula which works for all self-maps on all possible infra-solvmanifolds and which reduces to the old formulas in the case of infra-nilmanifolds or infra-solvmanifolds of type (R). Moreover, when viewing an infra-solvmanifold as a polynomial manifold, we recall that any map is homotopic to a polynomial map and we show how our formula can be translated in terms of the Jacobian of that polynomial map.

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

References

  1. Anosov, D.V.: The Nielsen numbers of maps of nil-manifolds. Uspekhi. Mat. Nauk. 40(4(224)), 133–134 (1985). English transl.: Russian Math. Surv. 40(4), 149–150 (1985)

  2. Baues, O.: Infra-solvmanifolds and rigidity of subgroups in solvable linear algebraic groups. Topology 43(4), 903–924 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  3. Brown, R.F.: The Lefschetz Fixed Point Theorem. Scott, Foresman and Company, Northbrook (1971)

    MATH  Google Scholar 

  4. Dekimpe, K.: Determining the translation part of the fundamental group of an infra-solvmanifold of type (R). Math. Proc. Camb. Philos. Soc. 122, 515–524 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  5. Dekimpe, K., Igodt, P.: The structure and topological meaning of almost-torsion free groups. Commun. Algebra 22(7), 2547–2558 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  6. Dekimpe, K., Igodt, P.: Polycyclic-by-finite groups admit a bounded-degree polynomial structure. Invent. Math. 129(1), 121–140 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  7. Dekimpe, K., Igodt, P., Lee, K.B.: Polynomial structures for nilpotent groups. Trans. Am. Math. Soc. 348(1), 77–97 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  8. Dekimpe, K., Lee, K.B., Raymond, F.: Bieberbach theorems for solvable Lie groups. Asian J. Math. 5(3), 499–508 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  9. Dekimpe, K., Tertooy, S., Van den Bussche, I.: Reidemeister spectra for solvmanifolds in low dimensions. Topol. Methods Nonlinear Anal. 53(2), 575–601 (2019)

    MathSciNet  MATH  Google Scholar 

  10. Deré, J.: NIL-affine crystallographic actions of virtually polycyclic groups. Transform. Groups 26(4), 1217–1240 (2021)

  11. Fadell, E., Husseini, S.: On a theorem of Anosov on Nielsen numbers for nilmanifolds. In Nonlinear functional analysis and its applications (Maratea, 1985), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 17., pp. 47–53. Reidel, Dordrecht (1986)

  12. Fel’shtyn, A., Lee, J.B.: The Nielsen and Reidemeister numbers of maps on infra-solvmanifolds of type \((R)\). Topol. Appl. 181, 62–103 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  13. Fel’shtyn, A., Troitsky, E., Leonov, Y.: Twisted conjugacy classes in saturated weakly branch groups. Geom. Dedicata 134, 61–73 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gonçalves, D., Wong, P.: Twisted conjugacy classes in nilpotent groups. J. Reine Angew. Math. 633, 11–27 (2009)

    MathSciNet  MATH  Google Scholar 

  15. Gorbacevič, V.V.: Lattices in Lie groups of type \((E)\) and \((R)\). Vestnik Moskov. Univ. Ser. I Mat. Meh. 30(6), 56–63 (1975)

    MathSciNet  MATH  Google Scholar 

  16. Ha, K.Y., Lee, J.B.: Averaging formula for Nielsen numbers of maps on infra-solvmanifolds of type (R)–corrigendum. Nagoya Math. J. 221(1), 207–212 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  17. Ha, K.Y., Lee, J.B., Penninckx, P.: Formulas for the Reidemeister, Lefschetz and Nielsen coincidence number of maps between infra-nilmanifolds. Fixed Point Theory Appl. 2012(39), 23 (2012)

    MathSciNet  MATH  Google Scholar 

  18. Hall, M. Jr.: The Theory of Groups. Chelsea Publishing Co., New York (1976). Reprinting of the 1968 edition

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

    MATH  Google Scholar 

  20. Jezierski, J., Marzantowicz, W.: Homotopy Methods in Topological Fixed and Periodic Point Theory, Topological Fixed Point Theory and Its Applications, vol. 3. Springer, Berlin (2006)

    Book  MATH  Google Scholar 

  21. Jiang, B.: Nielsen Fixed Point Theory, Contemp. Math., vol. 14. American Mathematical Society, Providence (1983)

    Book  Google Scholar 

  22. Keppelmann, E.C., McCord, C.K.: The Anosov theorem for exponential solvmanifolds. Pac. J. Math. 170(1), 143–159 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  23. Kim, S.W., Lee, J.B., Lee, K.B.: Averaging formula for Nielsen numbers. Nagoya Math. J. 178, 37–53 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  24. Kuroki, S., Yu, L.: On the equivalence of several definitions of compact infra-solvmanifolds. Proc. Jpn. Acad. Ser. A Math. Sci. 89(9), 114–118 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  25. Lee, J.B.: Nielsen spectrum of maps on infra-solvmanifolds modeled on \({\rm Sol}_0^4\). Bull. Korean Math. Soc. 57(4), 909–919 (2020)

    MathSciNet  MATH  Google Scholar 

  26. Lee, J.B., Lee, K.B.: Averaging formula for Nielsen numbers of maps on infra-solvmanifolds of type (R). Nagoya Math. J. 196, 117–134 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  27. Wang, H.-C.: Discrete subgroups of solvable Lie groups. I. Ann. Math. 2(64), 1–19 (1956)

    Article  MathSciNet  MATH  Google Scholar 

  28. Wecken, F.: Fixpunktklassen. III. Mindestzahlen von Fixpunkten. Math. Ann. 118, 544–577 (1942)

    Article  MathSciNet  MATH  Google Scholar 

  29. Wilking, B.: Rigidity of group actions on solvable lie groups. Math. Ann. 317, 195–237 (2000)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Karel Dekimpe.

Additional information

Publisher's Note

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

Research supported by long term structural funding—Methusalem grant of the Flemish Government.

Appendix

Appendix

We prove:

Lemma 5.7. Let \(X\in \mathbb {Z}^{n\times n}\) and \(\Phi \in \mathbb {Q}^{m\times m}\) be matrices and let \(A:\mathbb {Z}^m\rightarrow {\text {SL}}_n(\mathbb {Z})\) be an endomorphism, such that A(v) is net for all \(v\in \mathbb {Z}^m\). Suppose that \(\Phi \) does not have 1 as an eigenvalue, and that there exists \(k\in \mathbb {N}\), such that \(\Phi (k\mathbb {Z}^m)\subseteq \mathbb {Z}^m\) and \(XA(kv)=A(\Phi (kv))X\) for all \(v\in \mathbb {Z}^m\). Then, \(\textrm{det}(I-A(v)X)=\textrm{det}(I-X)\) for all \(v\in \mathbb {Z}^m\).

Proof

We follow the matrix analysis carried out by Keppelmann and McCord [22, Section 4]. This analysis consist of three steps.

Step 1. Reduction to the unipotent and semisimple case.

For \(v\in \mathbb {Z}^m\), write \(A(v)=U(v)T(v)\) with U(v) unipotent, T(v) semisimple and \([U(v),T(v)]=1\). This defines morphisms U, \(T:\mathbb {Z}^m\rightarrow \textrm{GL}_n(\mathbb {Q})\) and \([U(v),T(w)]=1\) for \(v\ne w\in \mathbb {Z}^m\), too. We show that it is sufficient to prove the lemma for \(A=U\) (the unipotent case) and \(A=T\) (the semisimple case).

Following [22], we say that X almost \(\Phi \)- commutes with A if there exists \(k\in \mathbb {N}\) such that \(XA(kv)=A(\Phi (kv))X\) for all \(v\in \mathbb {Z}^m\).

Observation: X almost \(\Phi \)- commutes with both U and T.

Set \(B(v):=\left( {\begin{matrix} A(kv)&{}0\\ 0&{}A (\Phi (kv)) \end{matrix}}\right) \) for all v in \(\mathbb {Z}^m\). Then

$$\begin{aligned} B_u(v):=\left( {\begin{matrix} U(kv)&{}0\\ 0&{}U (\Phi (kv)) \end{matrix}}\right) \quad \text {and} \quad B_s(v):=\left( {\begin{matrix} T(kv)&{}0\\ 0&{}T (\Phi (kv)) \end{matrix}}\right) \end{aligned}$$

are the unipotent and semisimple part of B(v), respectively. Consider the subgroup \(B:=\{B(v)\mid v\in \mathbb {Z}^m\}\) of \({\text {GL}}_{2n}(\mathbb {C})\). The relation \(XA(kv)=A(\Phi (kv))X\) is polynomial in the coefficients of B(v). As \(B_u(v)\) and \(B_s(v)\) are contained in the Zariski closure of B, they too must satisfy this relation. So X almost \(\Phi \)- commutes with U and with T. \(\square \)

Suppose now that

$$\begin{aligned} \begin{aligned}&(*)\ \textrm{det}(I-U(v)M)=\textrm{det}(I-M)\text { for every } M \text { almost }\Phi - \text {commuting with } U;\\&(\star )\ \textrm{det}(I-T(v)M)=\textrm{det}(I-M) \text { for every }M \text { almost }\Phi - \text {commuting with }T. \end{aligned} \end{aligned}$$

Say k satisfies \(XU(kv)=U(\Phi (kv))X\) for all \(v\in \mathbb {Z}^n\). Then, for all \(v, w\in \mathbb {Z}^n\)

$$\begin{aligned} T(v)XU(kw)&=T(v)U(\Phi (kw))X\\&=U(\Phi (kw))T(v)X, \end{aligned}$$

so in particular, T(v)X almost \(\Phi \)- commutes with U. Hence

$$\begin{aligned} \textrm{det}(I-A(v)X)&=\textrm{det}(I-U(v)T(v)X)\\&\overset{(*)}{=}\textrm{det}(I-T(v)X)\\&\overset{(**)}{=}\textrm{det}(I-X). \end{aligned}$$

We prove (\(*\)) in step 2 and (\(**\)) in step 3.

Step 2. The unipotent case.

Suppose that \(M\in M_n(\mathbb {C})\) almost \(\Phi \)- commutes with U, say \(d\in \mathbb {N}\) satisfies \(MU(dv)=U(\Phi (dv))M\) for all \(v\in \mathbb {Z}^m\). From the original version of this lemma, [22, Theorem 4.2], we already know that

$$\begin{aligned}&\textrm{det}(I-U(d\,v)M)=\textrm{det}(I-M) \end{aligned}$$

for all \(v\in \mathbb {Z}^m\). Fix \(x\in \mathbb {Z}^m.\) Then, for all \(z\in \mathbb {Z},\) also

$$\begin{aligned} \textrm{det}(I-U(dz\, x)M)=\textrm{det}(I-M). \end{aligned}$$

However, as U(x) is unipotent, the entries of \(U(t\, x)=U(x)^{t}\), \(t\in \mathbb {Z}\), are polynomials in t (depending on the entries of U(x), of course). Hence

$$\begin{aligned} \textrm{det}(I-U(t\, x)M)-\textrm{det}(I-M)\in \mathbb {Q}[t] \end{aligned}$$

is a polynomial vanishing on \(d\mathbb {Z}\), so it must be zero. We conclude that \(\textrm{det}(I-U(x)M)=\textrm{det}(I-M)\), as required.

Step 3. The semisimple case.

Suppose that \(M\in M_n(\mathbb {C})\) almost \(\Phi \)- commutes with T; we are to show that \(\textrm{det}(I-T(v)M)=\textrm{det}(I-M)\) for all \(v\in \mathbb {Z}^m\).

Choose a basis of common eigenvectors \(\{f_j\}_{j=1,\ldots , n}\) of the T(v)’s. Let \([x_{ij}]\in M_n(\mathbb {C})\) represent M with respect to the basis \(\{f_j\}\), and let \(\lambda _j(v)\) be the eigenvalue of T(v) associated to \(f_j\). Then, in this notation

$$\begin{aligned} \textrm{det}(I-T(v)M)=\sum _{\sigma \in \mathcal {S}_n}\left( {\text {sgn}}(\sigma )\prod _{i=1}^n \, \delta _{i\sigma (i)}-\lambda _i(v)\, x_{i\sigma (i)}\right) \end{aligned}$$

denoting \(\delta _{ij}\) the Kronecker delta. It is, therefore, sufficient to prove the following:

Reduction 1. \(\forall \sigma \in \mathcal {S}_n: \prod _{i=1}^n \, \delta _{i\sigma (i)}-\lambda _i(v) \, x_{i\sigma (i)}=\prod _{i=1}^n \, \delta _{i\sigma (i)}-x_{i\sigma (i)}\)

Take \(\sigma \in \mathcal {S}_n.\) Write \(\sigma =\sigma _1\circ \cdots \circ \sigma _l\) as a product of disjoint cycles \(\sigma _i\). For each element in \({\text {fix}}(\sigma )\), we add a ‘cycle’ of length 1. Formally, denoting \({\text {fix}}(\sigma )=\{e_1,\ldots , e_d\}\) with \(d=\#{\text {fix}}(\sigma )\), we set \(\sigma _{l+j}:=(e_j)\) for \(j\in \{1,\ldots , d\}\).

We can now partition the set \(\{1,\ldots , n\}\) according to these cycles: define

$$\begin{aligned} V(\sigma _j):= {\left\{ \begin{array}{ll} \{1,\ldots ,n\}\setminus {\text {fix}}(\sigma _j) &{}\text { if }j\le l,\\ \{e_j\} &{}\text { if }j>l. \end{array}\right. } \end{aligned}$$

Setting \(r=d+l\), we see that \(\{1,\ldots , n\}=\cup _{j=1}^r V(\sigma _j)\), so we can further reduce to

Reduction 2. \(\forall j\in \{1,\ldots r\}: \displaystyle \prod _{i\in V(\sigma _j)} \delta _{i\sigma _j(i)}-\lambda _i(v) \, x_{i\sigma _j(i)}=\displaystyle \prod _{i\in V(\sigma _j)} \delta _{i\sigma _j(i)}-x_{i\sigma _j(i)}\)

Take \(j\in \{1,\ldots , r\}\). Write

$$\begin{aligned} \sigma _j=\left( h\ \sigma _j(h)\ \sigma _j^2(h)\ \ldots \ \sigma _j^{s-1}(h)\right) . \end{aligned}$$

To shorten notation, we write \(\sigma ^i:=\sigma _j^i(h)\). So \(\sigma ^{s+1}=\sigma ^1\). We have to show that

figure k

We examine this statement more closely by distinguishing the cases \(s=1\) and \(s>1\).

  • If \(s=1\), statement (\(\circ \)) reads \(1-\lambda _h(v)\,x_{hh}=1-x_{hh}\), or equivalently, \(x_{hh}=0\) or \(\lambda _h(v)=1\).

  • If \(s>1\), statement (\(\circ \)) reads \(\prod _{i=1}^s\lambda _{\sigma ^i}(v)\,x_{\sigma ^i\sigma ^{i+1}}=\prod _{i=1}^s x_{\sigma ^i\sigma ^{i+1}}\), or equivalently, \(\prod _{i=1}^s x_{\sigma ^i\sigma ^{i+1}}=0\) or \(\prod _{i=1}^s\lambda _{\sigma ^i}(v)=1\).

So in both cases, statement (\(\circ \)) is equivalent to \(\prod _{i=1}^s x_{\sigma ^i\sigma ^{i+1}}=0\) or \(\prod _{i=1}^s\lambda _{\sigma ^i}(v)=1\). We will show the following:

Reduction 3. If \(\prod _{i=1}^s x_{\sigma ^i\sigma ^{i+1}}\ne 0\), then \(\prod _{i=1}^s\lambda _{\sigma ^i}(v)=1\).

Therefore, assume that \(\prod _{i=1}^s x_{\sigma ^i\sigma ^{i+1}}\ne 0\). As M almost \(\Phi \)- commutes with T, there exists \(k\in \mathbb {N}\) such that \(MT(kw)=T(\Phi (kw))M\) for every \(w\in \mathbb {Z}^m\). Take \(i\in \{1,\ldots , s\}\). Then

$$\begin{aligned} MT(kw)f_{\sigma ^{i+1}}&=M\lambda _{\sigma ^{i+1}}(kw)f_{\sigma ^{i+1}}\\&=\sum _{j=1}^n \lambda _{\sigma ^{i+1}}(kw)x_{j\sigma ^{i+1}}f_j\\ \text {and} T(\Phi (kw))Mf_{\sigma ^{i+1}}&=T(\Phi (kw))\left( \sum _{j=1}^nx_{j\sigma ^{i+1}}f_j\right) \\&=\sum _{j=1}^n x_{j\sigma ^{i+1}} T(\Phi (kw))(f_j)\\&=\sum _{j=1}^n x_{j\sigma ^{i+1}} \lambda _j(\Phi (kw))f_j. \end{aligned}$$

Equating \(MT(kw)=T(\Phi (kw))M\), we see that \( x_{j\sigma ^{i+1}}\ne 0\) implies \(\lambda _{\sigma ^{i+1}}(kw)=\lambda _j(\Phi (kw))\) for all \(j\in \{1,\ldots , n\}\). As \(x_{\sigma ^i\sigma ^{i+1}}\ne 0\) by assumption

figure l

for all \(i\in \{1,\ldots , s\}\) and \(w\in \mathbb {Z}^m\).

Take a basis \(\{e_1,\ldots , e_m\}\) of \(\mathbb {Z}^m\) such that \(\{e_1,\ldots , e_q\}\) (viewed as subset of \(\mathbb {R}^m)\) spans \(\ker ((\Phi ^T)^s-I)\) for some \(q\in \{0,\ldots , m\}\). Write \(\Phi =[\phi _{ij}]\) with respect to the basis \(\{e_j\}\), and write \(\lambda _{i,j}:=\lambda _{\sigma ^i}(e_j)\). (So again \(\lambda _{s+1,j}=\lambda _{1,j}\).)

We have to show that \(\prod _{i=1}^s\lambda _{\sigma ^i}(v)=1\) for every \(v\in \mathbb {Z}^m\). If \(v=\sum _{j=1}^m \alpha _j e_j\), then \(T(v)=\prod _{j=1}^m T(e_j)^{\alpha _j}\), hence \(\lambda _{\sigma ^i}(v)=\prod _{j=1}^m \lambda _{i,j}^{\alpha _j}\) for every \(i\in \{1,\ldots , s\}\). Therefore, it is sufficient to prove that

Reduction 4. \(\prod _{i=1}^s \lambda _{i,j}=1\) for all \(j\in \{1,\ldots ,m\}\).

We know exploit condition (\(\bullet \)). Thereto, take \(j\in \{1,\ldots , m\}\) and choose \(r_{i,j}\), \(\theta _{i,j}\in \mathbb {R}\), \(i\in \{1,\ldots , s\}\), satisfying \(\lambda _{i,j}=e^{r_{i,j}+2\pi i \,\theta _{i,j}}\). In this notation

$$\begin{aligned} \lambda _{\sigma ^{i+1}}(ke_j)&={\lambda _{i+1, j}}^k\\&=e^{kr_{i+1,j}}e^{2\pi i\, k\theta _{i+1,j}} \end{aligned}$$

when we agree that \(r_{s+1,j}:=r_{1,j}\) and \(\theta _{s+1,j}:=\theta _{1,j}\). Furthermore, since \(\Phi (ke_j)=\sum _{\alpha =1}^m k\phi _{\alpha j} e_\alpha ,\)

$$\begin{aligned} \lambda _{\sigma ^i}(\Phi (ke_j))&=\prod _{\alpha =1}^m\lambda _{i,\alpha }^{k\phi _{\alpha j}}\\&=\prod _{\alpha =1}^m e^{k\phi _{\alpha j}r_{i,\alpha }+2\pi i \, k\phi _{\alpha j}\theta _{i, \alpha }}\\&=e^{k\sum _{\alpha =1}^m \phi _{\alpha j }r_{i,\alpha }} e^{2\pi i\, k \sum _{\alpha =1}^m\phi _{\alpha j}\theta _{i,\alpha }}. \end{aligned}$$

Imposing condition (\(\bullet \)) implies that for all \(i\in \{1,\ldots , s\}\)

$$\begin{aligned} kr_{i+1, j}&=k\sum _{\alpha =1}^m \phi _{\alpha j }r_{i,\alpha }{} & {} \text {and}&k\theta _{i+1, j}&\equiv k\sum _{\alpha =1}^m \phi _{\alpha j }\theta _{i,\alpha } \mod \mathbb {Z}. \end{aligned}$$

Define \(R_i, \Theta _i\in \mathbb {R}^m \) as the vectors with j-th component equal to \(r_{i,j}\) and \(\theta _{i,j},\) respectively. Then

$$\begin{aligned} R_{i+1}&=\Phi ^T R_i{} & {} \text {and}&\Theta _{i+1}&\equiv \Phi ^T\Theta _i \mod \mathbb {Q}^m. \end{aligned}$$

Note that \(R_{s+1}=R_1\) and \(\Theta _{s+1}=\Theta _1\). Therefore, the above implies that

  • \(\Phi ^T\left( \sum _{i=1}^s R_i\right) =\sum _{i=1}^s R_i\);

  • \({\Phi ^T}^s(R_i)=R_i\);

  • \({\Phi ^T}^s(\Theta _i)\equiv \Theta _i \mod \mathbb {Q}^m\).

It follows from the first item that \(\sum _{i=1}^s R_i=0\) for \(\Phi ^{T}\) (and \(\Phi \)) does not have eigenvalue 1.

The second item implies that \(R_i\in \ker ({\Phi ^T}^s-I)\), hence \(r_{i,j}=0\) if \(j>q\). As \(\lambda _{i,j}\) cannot be a nontrivial root of unity, \(\theta _{i,j}\) must either be 0 or irrational if \(j>q\). In fact, \(\theta _{i,j}=0\) as \(({\Phi ^T}^s-I)(\Theta _i)\) must lie inside \(\mathbb {Q}^m\) and

$$\begin{aligned} {\Phi ^T}^s-I= \begin{pmatrix} 0 &{}\quad *\\ 0 &{}\quad \Phi ' \end{pmatrix} \end{aligned}$$

for some \(\Phi '\in \textrm{GL}_{m-q}(\mathbb {Q})\) (with respect to the basis \(\{e_j\}\)). So \({\Phi ^T}^s(\Theta _i)=\Theta _i\) as well.

However, then \(\Phi ^T\) fixes \(\sum _{i=0}^{s-1}{\Phi ^T}^i(\Theta _1)\), implying \(\sum _{i=0}^{s-1}{\Phi ^T}^i(\Theta _1)=0\) for \(\Phi \) does not have 1 as an eigenvalue. As \(\Theta _{i+1}\equiv {\Phi ^T}^i(\Theta _1) \mod \mathbb {Q}^m\), we conclude that \(\sum _{i=1}^s \Theta _i\in \mathbb {Q}^m\).

Translating \(\sum _{i=1}^s R_i=0\) and \(\sum _{i=1}^s \Theta _i\in \mathbb {Q}^m\) back to the \(r_{i,j}\)/\(\theta _{i,j}\)-notation gives

$$\begin{aligned} \sum _{i=1}^s r_{i,j}=0 \quad \text {and}\quad \sum _{i=1}^s \theta _{i,j}\in \mathbb {Q}\end{aligned}$$

for all \(j\in \{1,\ldots , m\}\). Hence

$$\begin{aligned} \prod _{i=1}^s \lambda _{i,j}&=\prod _{i=1}^s e^{r_{i,j}+2\pi i\, \theta _{i,j}}\\&=e^{\sum _{i=1}^s r_{i,j} + 2\pi i\sum _{i=1}^s \theta _{i,j}}\\&\in e^{2\pi i \,\mathbb {Q}} \end{aligned}$$

is a root of unity. As \(T(e_j)\) is net, this implies that \(\prod _{i=1}^s \lambda _{i,j}=1\), concluding the proof. \(\square \)

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

Dekimpe, K., van Den Bussche, I. An averaging formula for Nielsen numbers on infra-solvmanifolds. J. Fixed Point Theory Appl. 25, 20 (2023). https://doi.org/10.1007/s11784-022-01003-1

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11784-022-01003-1

Mathematics Subject Classification

Navigation