Skip to main content
Log in

On the peripheral point spectrum and the asymptotic behavior of irreducible semigroups of Harris operators

  • Published:
Positivity Aims and scope Submit manuscript

Abstract

Given a positive, irreducible and bounded \(C_0\)-semigroup on a Banach lattice with order continuous norm, we prove that the peripheral point spectrum of its generator is trivial whenever one of its operators dominates a non-trivial compact or kernel operator. For a discrete semigroup, i.e. for powers of a single operator \(T\), we show that the point spectrum of some power \(T^k\) intersects the unit circle at most in \(1\). As a consequence, we obtain a sufficient condition for strong convergence of the \(C_0\)-semigroup and for a subsequence of the powers of \(T\), respectively.

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. Aliprantis, C.D., Burkinshaw, O.: Positive operators. In: Pure and Applied Mathematics, vol 119. Academic Press Inc., Orlando (1985)

  2. Arendt, W.: Über das Spektrum regulärer Operatoren. PhD thesis, Eberhard-Karls-Universität zu Tübingen (1979)

  3. Arendt, W.: On the o-spectrum of regular operators and the spectrum of measures. Math. Z. 178(2), 271–287 (1981)

    Google Scholar 

  4. Arendt, W.: Positive semigroups of kernel operators. Positivity 12(1), 25–44 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  5. Arendt, W., Batty, C.J.K.: Tauberian theorems and stability of one-parameter semigroups. Trans. Am. Math. Soc. 306(2), 837–852 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  6. Arendt, W., Batty, C.J.K., Hieber, M., Neubrander, F.: Vector-valued Laplace transforms and Cauchy problems. Monographs in Mathematics, vol. 96. Birkhäuser, Basel (2001)

  7. Axmann, D.: Struktur-und Ergodentheorie irreduzibler Operatoren auf Banachverbänden. PhD thesis, Eberhard-Karls-Universität zu Tübingen (1980)

  8. Bogachev, V.I.: Measure Theory, vol. I. II. Springer, Berlin (2007)

  9. Davies, E.B.: Triviality of the peripheral point spectrum. J. Evol. Equ. 5(3), 407–415 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  10. Émilion, R.: Mean-bounded operators and mean ergodic theorems. J. Funct. Anal. 61(1), 1–14 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  11. Engel, K., Nagel, R.: One-parameter semigroups for linear evolution equations. In: Graduate Texts in Mathematics, vol. 194. Springer, New York (2000)

  12. Greiner, G.: Zur Perron-Frobenius-Theorie stark stetiger Halbgruppen. Math. Z. 177(3), 401–423 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  13. Greiner, G.: Spektrum und Asymptotik stark stetiger Halbgruppen positiver Operatoren. Sitzungsber. Heidelb. Akad. Wiss. Math.-Natur. Kl. pp. 55–80 (1982)

  14. Keicher, V.: On the peripheral spectrum of bounded positive semigroups on atomic Banach lattices. Arch. Math. (Basel) 87(4), 359–367 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  15. Krengel, U.: Ergodic theorems. In: de Gruyter Studies in Mathematics, vol 6. Walter de Gruyter& Co., Berlin (1985)

  16. Meyer-Nieberg, P.: Banach lattices. In: Universitext. Springer, Berlin (1991)

  17. Schaefer, H.H.: Banach lattices and positive operators. In: Die Grundlehren der mathematischen Wissenschaften, Band 215. Springer, New York (1974)

  18. Varopoulos, N.T.: Sets of multiplicity in locally compact abelian groups. Ann. Inst. Fourier (Grenoble). 16(fasc. 2):123–158 (1966)

    Google Scholar 

  19. Wolff, M.P.H.: Triviality of the peripheral point spectrum of positive semigroups on atomic Banach lattices. Positivity 12(1), 185–192 (2008)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The author was supported by the graduate school Mathematical Analysis of Evolution, Information and Complexity during the work on this article and he would like to thank Wolfgang Arendt for many helpful discussions and the anonymous referee for his/her constructive comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Moritz Gerlach.

Appendices

Appendix: Greiner’s zero-two law

We present the proof of Greiner’s zero-two law from [13] in a reformulation for Banach lattices with order continuous norm and without any continuity condition on the semigroup.

Throughout, let \(\fancyscript{T}=(T(t))_{t\in R}\) be a positive and bounded semigroup on \(E\), a Banach lattice with order continuous norm, and fix \(\tau >0\).

Theorem 5.1

(Greiner’s zero-two law) Assume that \(\mathrm {Fix} (\fancyscript{T})\) contains a quasi-interior point \(e\) of \(E_+\) and that there exists a strictly positive element in \(\mathrm {Fix} (\fancyscript{T}^{\prime })\). Then

$$\begin{aligned} E_2 \mathrel {\mathop :}=\{ y\in E : (T(t)\wedge T(t+\tau ))\vert y\vert = 0\quad \text{ for} \text{ all} t\in R \} \end{aligned}$$

and \(E_0\mathrel {\mathop :}=E_2^\bot \) are \(\fancyscript{T}\)-invariant bands. Moreover, if \(P\) denotes the band projection onto \(E_2\), then

$$\begin{aligned} \vert T(t) - T(t+\tau ) \vert Pe = 2Pe\quad for all t\in R \end{aligned}$$

and

$$\begin{aligned} \lim _{t\rightarrow \infty } \vert T(t) - T(t+\tau ) \vert (I-P)e =0. \end{aligned}$$

To simplify notation, for \(t \in R\) we define the positive operators

$$\begin{aligned} S(t)\mathrel {\mathop :}=T(t)\wedge T(t+\tau ) \quad \text{ and} \quad D(t) \mathrel {\mathop :}=\vert T(t) - T(t+\tau ) \vert \end{aligned}$$

on \(E\). It follows immediately that

$$\begin{aligned} S(t)x+\frac{1}{2}D(t)x = x \end{aligned}$$
(5.1)

for all \(x\in \mathrm {Fix} (\fancyscript{T})\) and \(t\in R\). Further properties of \(S(t)\) and \(D(t)\) are provided by the following lemma.

Lemma 5.2

Let \(x^{\prime } \in \mathrm {Fix} (\fancyscript{T}^{\prime })\) be strictly positive and \(\tau \in R\). Then the following assertions hold.

  1. (a)

    \(D(t)T(s) \ge D(t+s)\) and \(T(s)D(t)\ge D(t+s)\) for all \(t,s\in R\). Moreover, if \(x\in \mathrm {Fix} (\fancyscript{T})\), then \(\lim _{t\rightarrow \infty } D(t)x \in \mathrm {Fix} (\fancyscript{T})\).

  2. (b)

    \(S(t)T(s) \le S(t+s)\) and \(T(s)S(t)\le S(t+s)\) for all \(t,s\in R\). Moreover, if \(x\in \mathrm {Fix} (\fancyscript{T})\), then \(\lim _{t\rightarrow \infty } S(t)x \in \mathrm {Fix} (\fancyscript{T})\).

  3. (c)

    If \(\lim _{t\rightarrow \infty } S(t)x > 0\) for all \(0<x\in \mathrm {Fix} (\fancyscript{T})\), then \(\lim _{t\rightarrow \infty } S(t)^m x > 0\) for all \(m\in \mathbb{N }\) and \(x\in \mathrm {Fix} (\fancyscript{T})\), \(x>0\).

 

 

Proof

(a) For all \(t, s\in R\) one observes that

$$\begin{aligned} D(t)T(s) = \vert T(t)-T(t+\tau )\vert \cdot \vert T(s)\vert \ge \vert (T(t)-T(t+\tau ))T(s)\vert = D(t+s) \end{aligned}$$

and similarly that \(T(s)D(t)\ge D(t+s)\). Let \(x\in \mathrm {Fix} (\fancyscript{T})\), \(x\ge 0\). Then

$$\begin{aligned} D(t)x = D(t)T(s)x \ge D(t+s)x \end{aligned}$$

for all \(t,s\in R\). Hence, by the order continuity of the norm, \(y\mathrel {\mathop :}=\lim D(t)x\) exists in \(E\) and

$$\begin{aligned} T(s)y = \lim _{t\rightarrow \infty } T(s)D(t)x \ge \lim _{t\rightarrow \infty } D(t+s)x = y \ge 0. \end{aligned}$$

Finally, we conclude from \(\langle x^{\prime },T(s)y - y\rangle = 0\) that \(y\in \mathrm {Fix} (\fancyscript{T})\) because \(x^{\prime }\) is strictly positive.

(b) For all \(t,s\in R\) one observes that

$$\begin{aligned} S(t)T(s)&= \frac{1}{2} (T(t+s)+T(t+\tau +s)-D(t)T(s) )\\&\le \frac{1}{2} (T(t+s)+T(t+\tau +s)-D(t+s)) = S(t+s) \end{aligned}$$

and similarly that \(T(s)S(t) \le S(t+s)\). Hence, \(S(t)x\) is increasing for all positive \(x\in \mathrm {Fix} (\fancyscript{T})\). Since \(0 \le S(t) \le \frac{1}{2}(T(t)+T(t+\tau ))\), we conclude as in the proof of part (a) that \(\lim S(t)x\) exists in \(\mathrm {Fix} (\fancyscript{T})\) for all \(x\in \mathrm {Fix} (\fancyscript{T})\).

(c) Let \(x\in \mathrm {Fix} (\fancyscript{T})\), \(x>0\), and define recursively \(x_k \mathrel {\mathop :}=\lim _{t\rightarrow \infty } S(t)x_{k-1}\) for all \(k\in \mathbb{N }\) where \(x_0 \mathrel {\mathop :}=x\). Then \(x_k\in \mathrm {Fix} (\fancyscript{T})\) by part (b) and \(x_k>0\) by assumption. It follows by induction that

$$\begin{aligned} S(t)^m x-x_m = \sum _{j=1}^m S(t)^{m-j}( S(t)x_{j-1} -x_j ) \end{aligned}$$

for all \(m\in \mathbb N \) and \(t\in R\). As \(\Vert S(t) \Vert \le \sup _{t\in R}\Vert T(t)\Vert \), we conclude that

$$\begin{aligned} \lim _{t\rightarrow \infty } S(t)^m x = x_m>0 \end{aligned}$$

for all \(m\in \mathbb{N }\). \(\square \)

The key for the proof of the zero-two law is the following combinatorial lemma.

Lemma 5.3

For every \(m\in \mathbb{N }\)

$$\begin{aligned} 2^{-m}\left( \sum _{j=1}^{m} \biggr \vert \genfrac(){0.0pt}{}{m}{j}-\genfrac(){0.0pt}{}{m}{j-1} \biggr \vert +2\right) \le \frac{2}{\sqrt{m}}. \end{aligned}$$

 

Proof

For \(k\in \mathbb{N }\) and \(m=2k-1\) we have

$$\begin{aligned} \sum _{j=1}^m \biggr \vert \genfrac(){0.0pt}{}{m}{j}-\genfrac(){0.0pt}{}{m}{j-1} \biggr \vert +2&= 2\sum _{j=1}^{k-1}\biggr (\genfrac(){0.0pt}{}{m}{j}-\genfrac(){0.0pt}{}{m}{j-1}\biggr )+2\\&= 2\genfrac(){0.0pt}{}{m}{k-1} = \genfrac(){0.0pt}{}{2k}{k} \\&= \sum _{j=1}^k \biggr (\genfrac(){0.0pt}{}{m+1}{j}-\genfrac(){0.0pt}{}{m+1}{j-1}\biggr )+1\\&= \frac{1}{2}\sum _{j=1}^{m+1}\biggr \vert \genfrac(){0.0pt}{}{m+1}{j}-\genfrac(){0.0pt}{}{m+1}{j-1} \biggr \vert +1. \end{aligned}$$

It follows from Stirling’s formula that the central binomial coefficient can be estimated by

$$\begin{aligned} \genfrac(){0.0pt}{}{2k}{k} \le \frac{2^{2k}}{\sqrt{2k}} \quad (k\in \mathbb{N }). \end{aligned}$$

Thus, we obtain that

$$\begin{aligned} 2^{-m}\left( \sum _{j=1}^m \biggr \vert \genfrac(){0.0pt}{}{m}{j}-\genfrac(){0.0pt}{}{m}{j-1}\biggr \vert +2\right)&= 2^{-(m+1)} \left(\sum _{j=1}^{m+1}\biggr \vert \genfrac(){0.0pt}{}{m+1}{j}-\genfrac(){0.0pt}{}{m+1}{j-1}\biggr \vert +2\right)\\&= 2^{-m} \genfrac(){0.0pt}{}{2k}{k} \le \frac{2}{\sqrt{m+1}} \le \frac{2}{\sqrt{m}}, \end{aligned}$$

which completes the proof. \(\square \)

 

Proof of Theorem

By Lemma 5.2 (b), we have

$$\begin{aligned} S(t)\vert T(s)y \vert \le S(t)T(s)\vert y\vert \le S(t+s)\vert y \vert = 0 \end{aligned}$$

for all \(y\in E_2\) and \(t, s\in R\), which shows that \(E_2\) is \(\fancyscript{T}\)-invariant. Let \(x^{\prime }\in \mathrm {Fix} (\fancyscript{T}^{\prime })\) be strictly positive. Since

$$\begin{aligned} 0 \le T(t)Pe = PT(t)Pe \le PT(t)e = Pe \end{aligned}$$

and \(\langle x^{\prime },Pe - T(t)Pe\rangle =0\) for all \(t\in R\), it follows that \(Pe \in \mathrm {Fix} (\fancyscript{T})\). Define \(e_0 \mathrel {\mathop :}=(I-P)e \in E_0\) and \(e_2 \mathrel {\mathop :}=Pe\in E_2\). As \(e_0 = e-e_2 \in \mathrm {Fix} (\fancyscript{T})\), we conclude that \(E_0\), which equals the closure of the principle ideal generated by \(e_0\), is \(\fancyscript{T}\)-invariant.

It follows immediately from (5.1) that \(D(t)e_2 =2 e_2\) for all \(t\in R\). Hence, it remains to show that \(\lim D(t) e_0 = 0\). For simplicity, we omit the index \(0\) and write \(E=E_0\), \(e=e_0\), \(T(t)=T(t)|_{E_0}\) and so on. Now, \(\lim S(t)y > 0\) for all \(y\in \mathrm {Fix} (\fancyscript{T})\), \(y>0\), by Lemma 5.2 (b) and the definition of \(E_0\).

Assume that \(h\mathrel {\mathop :}=\lim D(t)e >0\). As \(h\le 2e\), there exists \(m\in \mathbb{N }\) such that

$$\begin{aligned} k \mathrel {\mathop :}=\biggr ( h -\frac{2}{\sqrt{m}} e\biggr )^+>0. \end{aligned}$$

Moreover, \(k\in \mathrm {Fix} (\fancyscript{T})\) since the fixed space is a sublattice. Indeed, if \(y\in \mathrm {Fix} (\fancyscript{T})\), it follows from \(T(t)\vert y\vert \ge \vert T(t)y\vert = \vert y\vert \) and \(\langle x^{\prime },T(t)\vert y\vert - \vert y\vert \rangle = 0\) for all \(t\in R\) that \(\vert y\vert \in \mathrm {Fix} (\fancyscript{T})\) because \(x^{\prime }\) is strictly positive. Now, Lemma 5.2 (c) yields that \(S(t_0)^m k > 0\) for some \(t_0\in R\). Let \(t_1 \mathrel {\mathop :}=m(t_0 + \tau )\) and define the operator

$$\begin{aligned} U \mathrel {\mathop :}=T(t_1) - 2^{-m} S(t_0)^m \biggr (I+T(\tau )\biggr )^m. \end{aligned}$$
(5.2)

It follows from \(S(t_0)(I+T(\tau )) \le T(t_0+\tau )+T(t_0)T(\tau ) = 2T(t_0+\tau )\) that \(U\) is positive. Moreover,

$$\begin{aligned} T(n t_1) = U^n + R_n 2^{-m} \biggr ( I + T(\tau )\biggr )^m \end{aligned}$$
(5.3)

for all \(n\in \mathbb{N }\) where \(R_1 \mathrel {\mathop :}=S(t_0)^m\) and \(R_{n+1} \mathrel {\mathop :}=U^nR_1 + R_nT(t_1)\). We infer from \(e=U^ne+R_ne\) that \(0\le R_n e \le e\) and \(0\le U^n e \le e\) for all \(n\in \mathbb{N }\). Now, by Lemma 5.2 (a) and Lemma 5.3, we obtain that

$$\begin{aligned} h \le D(nt_1)e&= \vert T(nt_1)(I-T(\tau ))\vert e \\&\le 2U^n e + R_n2^{-m} \left|\sum _{j=0}^m \genfrac(){0.0pt}{}{m}{j} T^j(\tau )(I-T(\tau ))\right|e \\&= 2U^n e + R_n 2^{-m} \left|\sum _{j=0}^m \genfrac(){0.0pt}{}{m}{j}T^j(\tau ) - \sum _{j=1}^{m+1}\genfrac(){0.0pt}{}{m}{j-1}T^j(\tau )\right|e \\&= 2U^n e + R_n2^{-m} \left( 2e + \sum _{j=1}^{m} \biggr \vert \genfrac(){0.0pt}{}{m}{j}-\genfrac(){0.0pt}{}{m}{j-1}\biggr \vert e\right)\\&\le 2 U^ne+R_n\frac{2}{\sqrt{m}}e \le 2 U^ne+\frac{2}{\sqrt{m}}e \end{aligned}$$

for every \(n\in \mathbb{N }\). Let \(y\mathrel {\mathop :}=\lim _{n\rightarrow \infty } U^n e \ge 0\). Then \(h\le 2 y + \frac{2}{\sqrt{m}} e\) and hence

$$\begin{aligned} 0 < k = \biggr ( h -\frac{2}{\sqrt{m}} e\biggr )^+ \le 2y. \end{aligned}$$

Since \(y\) is a fixed point of \(U\), Eq. (5.3) yields that \(T(n t_1)y \ge y \ge 0\) and we conclude from \(\langle x^{\prime },T(nt_1)y - y\rangle =0\) that \(T(n t_1)y=y\) for every \(n\in \mathbb{N }\). By Eq. (5.2) we have

$$\begin{aligned} 0 = S(t_0)^m \biggr ( I+T(\tau ) \biggr )^m y \ge S(t_0)^my \ge 0. \end{aligned}$$

Therefore, \(0 < S(t_0)^m k \le 2S(t_0)^m y = 0\) which contradicts the preceded observation that \(S(t_0)^mk>0\). Hence, \(h=\lim D(t)e = 0\). \(\square \)

 

Appendix: Axmann’s theorem

We give a proof Axmann’s theorem from [7], Satz 3.5] stating that not all powers of an irreducible Harris operator can be disjoint. A proof of this for \(E=L^p\) can be found in [4], Sec. 6].

We start with a version for \(L\)-spaces and reduce the general case to it in what follows. Let us recall that a Banach lattice \(E\) is said to be a \(L\) -space if

$$\begin{aligned} ||x + y ||= ||x||+ ||y||\end{aligned}$$

holds for all \(x,y \in E_+\).

Proposition 6.1

Let \(T\) be a positive and irreducible operator on a \(L\)-space \(E\) such that \(T\not \in (E^{\prime }\otimes E)^\bot \). Then there is \(n\in \mathbb N \), \(n\ge 2\), such that \(T\wedge T^n >0\).

 

Proof

Aiming for a contradiction, we assume that \(T\wedge T^n =0\) for all \(n \ge 2\). Since \(E\) is a \(L\)-space, we may identify \(E^{\prime }\) with \(C(K)\) for some compact space \(K\) by Kakutani’s theorem [16], Thm. 2.1.3]. For \(n,m \in \mathbb N \), \(n\ge 2\), define

$$\begin{aligned} A_n \mathrel {\mathop :}=\{ T^{\prime }h + T^{\prime n}(\mathbb 1 -h) : h\in C(K),\, 0\le h\le \mathbb 1 \} \subseteq C(K) \end{aligned}$$

and

$$\begin{aligned} O_{n,m} \mathrel {\mathop :}=\{t \in K : h(t) < 1/m \text{ for} \text{ some} h\in A_n\}. \end{aligned}$$

It follows from our assumption and Synnatschke’s theorem [16], Prop. 1.4.17] that

$$\begin{aligned} \inf A_n = (T^{\prime } \wedge T^{\prime n})\mathbb 1 = (T\wedge T^n)^{\prime } \mathbb 1 = 0 \end{aligned}$$

for all \(n\ge 2\). Now, we show that each of the open sets \(O_{n,m}\) is dense in \(K\). Assume the opposite. Then there exists a non-empty open set \(U\subseteq K\backslash O_{n,m}\) for some \(n,m \in \mathbb{N }\). By Urysohn’s theorem, we can construct a continuous function \(g:K\rightarrow [0,\frac{1}{m}]\) vanishing on \(O_{n,m}\) such that \(g(t_0)=\frac{1}{m}\) for some \(t_0 \in U\). Hence, \(g>0\) is a lower bound of \(A_n\), which is impossible. Therefore, every \(O_{n,m}\) and, by Baire’s theorem, also \(G\mathrel {\mathop :}=\cap _{n,m} O_{n,m}\) is dense in \(K\). Note that for all \(t\in G\) and \(n\ge 2\) one has that

$$\begin{aligned} \langle T^{\prime \prime }\delta _t \wedge T^{\prime \prime n} \delta _t,\mathbb 1 \rangle = \inf \{ h(t) : h\in A_n \} = 0 \end{aligned}$$

and consequently \(T^{\prime \prime } \delta _t \wedge T^{\prime \prime n} \delta _t = 0\).

By assumption, there are \(x\in E_+\) and \(y^{\prime }\in E^{\prime }_+\) such that \(T\) is not disjoint from \(R = y^{\prime } \otimes x\). Then \(R^{\prime }= x\otimes y^{\prime }\) corresponds to a rank-one operator \(\mu \otimes g\) on \(C(K)\) for some \(\mu \in C(K)^{\prime }_+\) and \(g\in C(K)_+\). Since \(R^{\prime }\wedge T^{\prime } \ge (R\wedge T)^{\prime } >0\) there exists some \(e\in C(K)_+\), such that \(\varrho \mathrel {\mathop :}=(R^{\prime }\wedge T^{\prime })e>0\). For all \(0\le h\le e\) and \(t\in K\) it follows from

$$\begin{aligned} \varrho = (R^{\prime }\wedge T^{\prime })h + (R^{\prime }\wedge T^{\prime })(e-h) \le R^{\prime }h+T^{\prime }(e-h) \end{aligned}$$

that

$$\begin{aligned} \varrho (t) \le \langle R^{\prime }h,\delta _t\rangle +\langle e-h,T^{\prime \prime }\delta _t\rangle = g(t)\langle \mu ,h\rangle +\langle e-h,T^{\prime \prime }\delta _t\rangle . \end{aligned}$$

Taking the infimum over all \(0\le h\le e\) yields that \(\varrho (t) \le (g(t)\mu \wedge T^{\prime \prime }\delta _t)e\) for all \(t\in K\). Now, fix \(t\in \{ s\in K : \varrho (s)>0\} \cap G\) which exists since \(G\) is dense in \(K\). Then \(\nu \mathrel {\mathop :}=g(t)\mu \wedge T^{\prime \prime }\delta _t >0\) because \(\varrho (t)>0\). As \(\nu \) is dominated by \(g(t)\mu \) and \(\mu \) corresponds to \(x \in E\), \(\nu \) itself corresponds to a vector \(v\in E\), \(v>0\), since \(E\) is an ideal in \(E^{\prime \prime }\). This vector \(v\) satisfies

$$\begin{aligned} v \wedge T^n v \le T^{\prime \prime }\delta _t \wedge T^{\prime \prime n+1} \delta _t = 0 \end{aligned}$$

for all \(n\in \mathbb{N }\) because \(t\in G\).

Finally, we consider \(w\mathrel {\mathop :}=Tv\). If \(w=0\), then the closed ideal \(\overline{E_v}\) is \(T\)-invariant and non-trivial since \(T\ne 0\). If \(w>0\), then the closure of the \(T\)-invariant ideal

$$\begin{aligned} J\mathrel {\mathop :}=\{ z\in E : |z|\le c(w+Tw+\dots +T^k w) \text{ for} \text{ some} c>0 \text{ and} m\in \mathbb{N }\} \end{aligned}$$

is non-trivial because \(w\in \overline{J}\) and \(v \in J^\bot \). In both cases, \(T\) cannot be irreducible. Thus, we conclude that \(T\wedge T^n > 0\) for some \(n\ge 2\). \(\square \)

 

Theorem 6.2

Let \(T\) be a positive and irreducible operator on \(E\), a Banach lattice with order continuous norm, such that \(T\not \in (E^{\prime }\otimes E)^\bot \). Then there is \(n\in \mathbb{N }\), \(n\ge 2\), such that \(T\wedge T^n >0\).

 

Proof

Fix \(\lambda > ||T ||\) and \(y^{\prime }\in E^{\prime }_+\), \(y^{\prime }\ne 0\). Then \(z^{\prime } \mathrel {\mathop :}=(\lambda - T^{\prime })^{-1}y^{\prime }\) satisfies \(\lambda z^{\prime } - T^{\prime }z^{\prime } = y^{\prime }\) and hence \(T^{\prime }z^{\prime } \le \lambda z^{\prime }\). This implies that the closed ideal \(\{ x\in E : \langle |x|,z^{\prime }\rangle =0\}\) is \(T\)-invariant and thus equal to \(\{0\}\). Therefore, \(||x||_{z^{\prime }} \mathrel {\mathop :}=\langle z^{\prime },|x|\rangle \) defines an order continuous lattice norm on \(E\). Let \((F,||\cdot ||_F)\) be the completion of \((E,||\cdot ||_{z^{\prime }})\), i.e. the closure of \(E\) in \((E,||\cdot ||_{z^{\prime }})^{\prime \prime }\). Then \(F\) is an \(L\)-space and, since \(||Tx ||_F \le \lambda ||T||_F\) for all \(x\in E\), \(T\) uniquely extends to a positive operator \(\tilde{T}\) on \(F\).

Now, it follows as in the proof of Proposition 3.2 that \(\tilde{T}\) is irreducible and \(\tilde{T} \not \in (F^{\prime }\otimes F)^{\bot }\).

Since the norm on \(E\) is order continuous, so is \(z^{\prime }\) and hence \(E\) is an ideal in \(F\) by [17], Lem. IV 9.3]. Thus, for \(x\in E_+\) and \(n\in \mathbb{N }\) we observe that

$$\begin{aligned} (\tilde{T}\wedge \tilde{T}^n)x&= \inf _F \{ \tilde{T}(x-y) +\tilde{T}^n y : y\in F,\, 0\le y\le x\}\\&= \inf _E \{ T(x-y)+T^n y : y\in E,\, 0\le y\le x\} \\&= (T\wedge T^n)x. \end{aligned}$$

As \(E_+\) is dense in \(F_+\), this shows that \(\tilde{T} \wedge \tilde{T}^n = 0\) if and only if \(T \wedge T^n = 0\). Thus, the assertion follows from Proposition 6.1. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gerlach, M. On the peripheral point spectrum and the asymptotic behavior of irreducible semigroups of Harris operators. Positivity 17, 875–898 (2013). https://doi.org/10.1007/s11117-012-0210-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11117-012-0210-8

Keywords

Mathematics Subject Classification (2000)

Navigation