Skip to main content
Log in

Nakayama Algebras and Fuchsian Singularities

  • Research
  • Published:
Algebras and Representation Theory Aims and scope Submit manuscript

Abstract

This present paper is devoted to the study of a class of Nakayama algebras \(N_n(r)\) given by the path algebra of the equioriented quiver \(\mathbb {A}_n\) subject to the nilpotency degree r for each sequence of r consecutive arrows. We show that the Nakayama algebras \(N_n(r)\) for certain pairs (nr) can be realized as endomorphism algebras of tilting objects in the bounded derived category of coherent sheaves over a weighted projective line, or in its stable category of vector bundles. Moreover, we classify all the Nakayama algebras \(N_n(r)\) of Fuchsian type, that is, derived equivalent to the bounded derived categories of extended canonical algebras. We also provide a new way to prove the classification result on Nakayama algebras of piecewise hereditary type, which have been done by Happel–Seidel before.

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

Availability of data and materials

All data generated or analysed during this study are included in this article.

References

  1. Auslander, M., Reiten, I., Smalø, S.O.: Representation theory of Artin algebras. Cambridge Studies in Advanced Mathematics, vol. 36. Cambridge University Press, Cambridge (1995)

  2. Barot, M., Lenzing, H.: One-point extensions and derived equivalence. J. Algebra 264(1), 1–5 (2003)

    Article  MathSciNet  Google Scholar 

  3. Bondal, A.I., Kapranov M.M.: Representable functors, Serre functors, and reconstructions. Izv. Akad. Nauk SSSR Ser. Mat. 53(6), 1183–1205, 1337 (1989)

  4. Buchweitz, R.-O.: Maximal Cohen-Macaulay Modules and Tate-cohomology over Gorenstein rings. Unpublished manuscript, pp. 155, (1987)

  5. Crawley-Boevey, W.: Connections for weighted projective lines. J. Pure Appl. Algebra 215(1), 35–43 (2011)

    Article  MathSciNet  Google Scholar 

  6. Geigle, W., Lenzing, H.: A class of weighted projective curves arising in representation theory of finite-dimensional algebras. In Singularities, representation of algebras, and vector bundles (Lambrecht, 1985), volume 1273 of Lecture Notes in Math., pp. 265–297. Springer, Berlin, (1987)

  7. Geigle, W., Lenzing, H.: Perpendicular categories with applications to representations and sheaves. J. Algebra 144(2), 273–343 (1991)

    Article  MathSciNet  Google Scholar 

  8. Happel, D.: Triangulated categories in the representation theory of finite-dimensional algebras. London Mathematical Society Lecture Note Series, vol. 119. Cambridge University Press, Cambridge (1988)

  9. Happel, D.: The trace of the Coxeter matrix and Hochschild cohomology. Linear Algebra Appl. 258, 169–177 (1997)

    Article  MathSciNet  Google Scholar 

  10. Happel, D.: A characterization of hereditary categories with tilting object. Invent. Math. 144(2), 381–398 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  11. Happel, D., Seidel, U.: Piecewise hereditary Nakayama algebras. Algebr. Represent. Theory 13(6), 693–704 (2010)

    Article  MathSciNet  Google Scholar 

  12. Happel, D., Zacharia, D.: A homological characterization of piecewise hereditary algebras. Math. Z. 260(1), 177–185 (2008)

    Article  MathSciNet  Google Scholar 

  13. Happel, D., Zacharia, D.: Homological properties of piecewise hereditary algebras. J. Algebra 323(4), 1139–1154 (2010)

    Article  MathSciNet  Google Scholar 

  14. Hille, L., Müller, J.: On tensor products of path algebras of type \(A\). Linear Algebra Appl. 448, 222–244 (2014)

    Article  MathSciNet  Google Scholar 

  15. Krause, H.: The stable derived category of a Noetherian scheme. Compos. Math. 141(5), 1128–1162 (2005)

    Article  MathSciNet  Google Scholar 

  16. Kussin, D., Lenzing, H., Meltzer, H.: Triangle singularities, ADE-chains, and weighted projective lines. Adv. Math. 237, 194–251 (2013)

    Article  MathSciNet  Google Scholar 

  17. Lenzing, H.: Wild canonical algebras and rings of automorphic forms. In Finite-dimensional algebras and related topics (Ottawa, ON, 1992), volume 424 of NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., pp. 191–212. Kluwer Acad. Publ., Dordrecht, (1994)

  18. Lenzing, H.: Coxeter transformations associated with finite-dimensional algebras. In Computational methods for representations of groups and algebras (Essen, 1997), volume 173 of Progr. Math., pp. 287–308. Birkhäuser, Basel, (1999)

  19. Lenzing, H.: Weighted projective lines and applications. In Representations of algebras and related topics, EMS Ser. Congr. Rep., pp. 153–187. Eur. Math. Soc., Zürich, (2011)

  20. Lenzing, H.: The algebraic theory of fuchsian singularties. In Advances in representation theory of algebras, volume 761 of Contemp. Math., pp. 171–190. Amer. Math. Soc., [Providence], RI, [2021] (2021)

  21. Lenzing, H., de la Peña, J.A.: Wild canonical algebras. Math. Z. 224(3), 403–425 (1997)

    Article  MathSciNet  Google Scholar 

  22. Lenzing, H., de la Peña, J.A.: On the growth of the Coxeter transformations of derived-hereditary algebras. Int. J. Algebra 1(1–4), 95–112 (2007)

    Article  MathSciNet  Google Scholar 

  23. Lenzing, H., de la Peña, J.A.: Spectral analysis of finite dimensional algebras and singularities. In Trends in representation theory of algebras and related topics, EMS Ser. Congr. Rep., pp. 541–588. Eur. Math. Soc., Zürich, (2008)

  24. Lenzing, H., de la Peña, J.A.: Extended canonical algebras and Fuchsian singularities. Math. Z. 268(1–2), 143–167 (2011)

    Article  MathSciNet  Google Scholar 

  25. Lenzing, H., de la Peña, J.A.: A Chebysheff recursion formula for Coxeter polynomials. Linear Algebra Appl. 430(4), 947–956 (2009)

    Article  MathSciNet  Google Scholar 

  26. Milnor, J.: On the \(3\)-dimensional Brieskorn manifolds \(M(p,q,r)\). In Knots, groups, and 3-manifolds (Papers dedicated to the memory of R. H. Fox), pp. 175–225. Ann. of Math. Studies, No. 84. (1975)

  27. Neumann, W.D.: Brieskorn complete intersections and automorphic forms. Invent. Math. 42, 285–293 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  28. Orlov, D.: Derived categories of coherent sheaves and triangulated categories of singularities. In: Algebra, arithmetic, and geometry: in honor of Yu. I. Manin. Vol. II, volume 270 of Progr. Math., pp. 503–531. Birkhäuser Boston, Boston, MA, (2009)

  29. Ringel, C.M.: Tame algebras and integral quadratic forms. Lecture Notes in Mathematics, vol. 1099. Springer-Verlag, Berlin (1984)

  30. Ringel, C.M.: The spectral radius of the Coxeter transformations for a generalized Cartan matrix. Math. Ann. 300(2), 331–339 (1994)

    Article  MathSciNet  Google Scholar 

  31. Sato, M.: Periodic Coxeter matrices and their associated quadratic forms. Linear Algebra Appl. 406, 99–108 (2005)

    Article  MathSciNet  Google Scholar 

  32. Simson, D., Skowroński, A.: Elements of the representation theory of associative algebras. Vol. 3, volume 72 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge. Representation-infinite tilted algebras (2007)

Download references

Funding

This work was partially supported by the Natural Science Foundation of Xiamen (No. 3502Z20227184) and the Natural Science Foundation of Fujian Province (No. 2022J01034), and the National Natural Science Foundation of China (No. 12271448), and by the Alexander von Humboldt Foundation in the framework of the Alexander von Humboldt Professorship endowed by the German Federal Ministry of Education and Research.

Author information

Authors and Affiliations

Authors

Contributions

Helmut Lenzing, Hagen Meltzer and Shiquan Ruan wrote the main manuscript text. All authors reviewed the manuscript.

Corresponding author

Correspondence to Shiquan Ruan.

Ethics declarations

Ethical approval

Not applicable.

Competing interests

The authors declared that they have no conflicts of interest to this work.

Additional information

Presented by: Henning Krause

Publisher's Note

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

Appendix A: Extension-free property

Appendix A: Extension-free property

Let \(\mathcal {D}\) be a Hom-finite triangulated category with Serre functor \({\mathbb {S}}=\tau [1]\), where \(\tau \) is the Auslander-Reiten translation and [1] denotes the suspension functor of \(\mathcal {D}\). The Serre duality of \(\mathcal {D}\) is given by the following natural isomorphism

$$\begin{aligned} {\textrm{Hom}}_{\mathcal {D}}(X,Y)\cong D \mathrm{Hom}_{\mathcal {D}}(Y, {\mathbb {S}}X), \end{aligned}$$

which is functorial in \(X,Y\in \mathcal {D}\).

Lemma A.1

Let \(X\in \mathcal {D}\) and \(U=\bigoplus \limits _{k=0}^{m}{\mathbb {S}}^{k}X\) for some \(m\ge 0\). Then U is extension-free in \(\mathcal {D}\) if and only if \(\mathrm{{Hom}}_{\mathcal {D}}(X, {\mathbb {S}}^{k}X[n])=0\) for any \(1\le k\le m+1\) and any non-zero integer n.

Proof

By definition, U is extension-free if and only if for any \(n\ne 0\) and \(0\le i,j \le m\), \(\mathrm{{Hom}}_{\mathcal {D}} ({\mathbb {S}}^{i}X, {\mathbb {S}}^{j}X[n])=0\), or equivalently, \(\mathrm{{Hom}}_{\mathcal {D}} (X, {\mathbb {S}}^{k}X[n])=0\) for any \(n\ne 0\) and \(-m\le k \le m\). For \(-m\le k\le 0\), we have \(\mathrm{{Hom}}_{\mathcal {D}} (X, {\mathbb {S}}^{k}X[n])\cong D\mathrm{{Hom}}_{\mathcal {D}} (X, {\mathbb {S}}^{1-k}X[-n])\) by using Serre duality, where \(1\le 1-k\le m+1\). This finishes the proof. \(\square \)

Lemma A.2

Let \(X,Y\in \mathcal {D}\) and \(U=\bigoplus \limits _{k=0}^{m}{\mathbb {S}}^{k}X\), \(V=\bigoplus \limits _{k=0}^{r}{\mathbb {S}}^{k}Y\) for some \(m,r\ge 0\). Assume both of U and V are extension-free. Then \(U\oplus V\) is extension-free if and only if \(\mathrm{{Hom}}_{\mathcal {D}}(X, {\mathbb {S}}^{k}Y[n])=0\) for any \(-m\le k\le r+1\) and any non-zero integer n.

Proof

By definition, \(U\oplus V\) is extension-free if and only if for any \(n\ne 0\), \(0\le i \le m\) and \(0\le j \le r\), \(\mathrm{{Hom}}_{\mathcal {D}} ({\mathbb {S}}^{i}X, {\mathbb {S}}^{j}Y[n])=0\) and \(\mathrm{{Hom}}_{\mathcal {D}} ({\mathbb {S}}^{j}Y[n], {\mathbb {S}}^{i}X)=D\mathrm{{Hom}}_{\mathcal {D}} ({\mathbb {S}}^{i}X, {\mathbb {S}}^{j+1}Y[n])=0\). Equivalently, \(\mathrm{{Hom}}_{\mathcal {D}} (X, {\mathbb {S}}^{k}Y[n])=0\) for any \(n\ne 0\) and \(-m\le k \le r+1\). This proves the result. \(\square \)

In the following, we will show that the objects \(T_{(2,4,5)}\), \(T_{(2,4,7)}\), \(T_{(2,5,5)}\), \(T_{(2,5,6)}\) constructed in Theorem 5.5 are extension-free in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}\) case by case.

Recall that for any weight type \((p_1,p_2,p_3)\), there is a surjective group homomorphism \(\delta :\mathbb {L}(p_1,p_2,p_3)\rightarrow \mathbb {Z}\) given by \(\delta (\vec {x}_i)=\frac{p}{p_i}\) for \(1\le i\le 3\), where \(p=\mathrm{l.c.m.}(p_1,p_2,p_3)\). We denote the \(\delta \)-datum \((\delta (\vec {c}); \delta (\vec {x}_1), \delta (\vec {x}_2), \delta (\vec {x}_3); \delta (\vec {\omega }))\) by \(\delta (\vec {c}; \vec {x}_1, \vec {x}_2, \vec {x}_3; \vec {\omega })\) for convenience.

For a weighted projective line \(\mathbb {X}\) of negative Euler characteristic, the associated Fuchsian singularity R is the restricted subalgebra \(S|_{\mathbb {Z}\vec {\omega }}\) of the coordinate algebra of \(\mathbb {X}\) defined in Eq. 2.1, i.e.,

$$\begin{aligned} R=\bigoplus _{n\ge 0}\textrm{Hom}_{}({\mathcal O}_{\mathbb {X}}, \tau ^n{\mathcal O}_{\mathbb {X}})=S|_{\mathbb {Z}\vec {\omega }}. \end{aligned}$$

1.1 A.1 The Weight Type (2,4,5)

Assume \(\mathbb {X}\) has weight type (2,4,5). By easy calculation we get the \(\delta \)-datum \(\delta (\vec {c}; \vec {x}_1, \vec {x}_2, \vec {x}_3; \vec {\omega })\)\(=(20;10,5,4;1)\).

According to [24], the semigroup \(\{n\vec {\omega }| n\vec {\omega }\ge 0\}\) of \(\mathbb {Z}\vec {\omega }\) has minimal generating system \(\{4\vec {\omega }=\vec {x}_3; 10\vec {\omega }=2\vec {x}_2; 15\vec {\omega }=\vec {x}_1+\vec {x}_2\}\). Hence, the subalgebra \(R=S|_{\mathbb {Z}\vec {\omega }}\cong k[x,y,z]/(f)\), where \(x=x_3, y=x_2^2, z=x_1x_2\) and \(f=z^2+y^3+x^5y\). Moreover, R is \(\mathbb {Z}\)-graded in the sense that \(\deg (x,y,z;f)=(4,10,15;30)\). Thus the second suspension functor of \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,5)\) is given by degree shift: \([2]=(30\vec {\omega })=(\vec {c}+2\vec {x}_2)\).

Since \(4\vec {\omega }=\vec {x}_3>0\), the set \(\mathcal {S}\) defined in Eq. 5.1 has the form

$$\begin{aligned} \mathcal {S}=\{\vec {x}\, |\, 0\le \vec {x}\le n\vec {\omega }+\vec {c}, \text {\ for\ any \ } 2\le n\le 5\} =\{0, \vec {x}_2\}. \end{aligned}$$

Hence each line bundle belongs to the orbit \(\tau ^{\mathbb {Z}}{\mathcal O}\) or \(\tau ^{\mathbb {Z}}{\mathcal O}(\vec {x}_2)\) by Proposition 5.1.

By Proposition 5.3, the projective cover of \({\mathcal O}(\vec {x}_2)\), under \(\tau ^{\mathbb {Z}}{\mathcal O}\)-exact structure on \(\textrm{vect}-\mathbb {X}\), is given by \(\mathcal {P}({\mathcal O}(\vec {x}_2))={\mathcal O}\oplus {\mathcal O}(-5\vec {\omega })\), which fits into the following exact sequence:

(A.1)

Lemma A.3

In \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,5)\), we have \({\mathcal O}(\vec {x}_2)[1]={\mathcal O}(\vec {x}_2)(15\vec {\omega }).\)

Proof

By Eq. A.1 we have \({\mathcal O}(\vec {x}_2)[-1]={\mathcal O}(-\vec {x}_1)={\mathcal O}(\vec {x}_2)(-15\vec {\omega })\). Then the result follows from \([2]=(30\vec {\omega })\). \(\square \)

Moreover, by Proposition 5.4, we have

Lemma A.4

In \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,5)\), we have

$$\begin{aligned} \underline{\textrm{Hom}}({\mathcal O}(\vec {x}_2-\vec {x}), {\mathcal O}(\vec {x}_2))\ne 0 \quad \text {if\ and\ only\ if}\quad 0\le \vec {x}\le 4\vec {x}_3. \end{aligned}$$
(A.2)

Proposition A.5

\(T_{(2,4,5)}\) is extension-free in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,5)\).

Proof

Recall that \(T_{(2,4,5)}=\bigoplus \limits _{k=0}^{10}L(k\vec {x}_3)\), where \(L={\mathcal O}(\vec {x}_2)\). For construction we assume \(T_{(2,4,5)}\) is not extension-free, then there exist some integers \(0\le a, b \le 10\) and \(0\ne m\in \mathbb {Z}\), such that

$$\begin{aligned} \underline{\textrm{Hom}}(L(a\vec {x}_{3}), L(b\vec {x}_{3})[m])\ne 0. \end{aligned}$$

Note that \(\vec {x}_3=4\vec {\omega }\) and \(L[1]=L(15\vec {\omega })\). According to Eq. A.2, we have

$$\begin{aligned} (b-a)\vec {x}_{3}+15m\vec {\omega }=k\vec {x}_{3} \text {for\ some\ } 0\le k\le 4. \end{aligned}$$
(A.3)

It follows that \(4(b-a)+15m=4k\) since \(\vec {x}_3=4\vec {\omega }\). Hence 4|m, say, \(m=4r\) for some integer \(r\ne 0\). Then we have \(0\le k=b-a+15r\le 4\), a contradiction to the assumption \(0\le a, b \le 10\). We are done. \(\square \)

1.2 A.2 The Weight Type (2,4,7)

Assume \(\mathbb {X}\) has weight type (2,4,7). Then the \(\delta \)-datum is given by \(\delta (\vec {c}; \vec {x}_1, \vec {x}_2, \vec {x}_3; \vec {\omega })=(28;14,7,4;3)\).

According to [24], the semigroup \(\{n\vec {\omega }| n\vec {\omega }\ge 0\}\) of \(\mathbb {Z}\vec {\omega }\) has minimal generating system \(\{4\vec {\omega }=3\vec {x}_3; 6\vec {\omega }=2\vec {x}_2+\vec {x}_3; 7\vec {\omega }=\vec {x}_1+\vec {x}_2\}\). Hence, the subalgebra \(R=S|_{\mathbb {Z}\vec {\omega }}\cong k[x,y,z]/(f)\), where \(x=x_3^3, y=x_2^2x_3, z=x_1x_2\) and \(f=y^3+x^3y+xz^2\). Moreover, R is \(\mathbb {Z}\)-graded in the sense that \(\deg (x,y,z;f)=(4,6,7;18)\). Thus the second suspension functor of \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\) is given by degree shift: \([2]=(18\vec {\omega })=(\vec {c}+2\vec {x}_2+3\vec {x}_3)\).

Since \(4\vec {\omega }=3\vec {x}_3>0\), the set \(\mathcal {S}\) defined in Eq. 5.1 has the form

$$\begin{aligned} \mathcal {S}=\{\vec {x}\, |\, 0\le \vec {x}\le n\vec {\omega }+\vec {c}, \text {\ for\ any \ } 2\le n\le 5\} =\{\vec {x}\,|\, 0\le \vec {x}\le \vec {x}_2+2\vec {x}_3\}. \end{aligned}$$

Hence each line bundle belongs to the orbit \(\tau ^{\mathbb {Z}}{\mathcal O}(\vec {x})\) for some \(0\le \vec {x}\le \vec {x}_2+2\vec {x}_3\).

By Proposition 5.3, the projective covers of \({\mathcal O}(\vec {x}_2)\) and \({\mathcal O}(\vec {x}_2+\vec {x}_3)\), under \(\tau ^{\mathbb {Z}}{\mathcal O}\)-exact structure on \(\textrm{vect}-\mathbb {X}\), are given by \(\mathcal {P}({\mathcal O}(\vec {x}_2))={\mathcal O}\oplus {\mathcal O}(-5\vec {\omega })\) and \(\mathcal {P}({\mathcal O}(\vec {x}_2+\vec {x}_3))={\mathcal O}\oplus {\mathcal O}(-\vec {\omega })\), which fit into the following exact sequences:

(A.4)
(A.5)

Lemma A.6

The following statements hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\).

  1. (1)

    \({\mathcal O}(\vec {x}_2)[1]={\mathcal O}(\vec {x}_2+\vec {x}_3)(7\vec {\omega });\)

  2. (2)

    \({\mathcal O}(\vec {x}_2+\vec {x}_3)[1]={\mathcal O}(\vec {x}_2)(11\vec {\omega }).\)

Proof

Since we have exact sequences (A.4) and (A.5) for projective covers, the following hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\):

  • \({\mathcal O}(\vec {x}_2)[-1]={\mathcal O}(-\vec {x}_1-2\vec {x}_3)={\mathcal O}(\vec {x}_2+\vec {x}_3)(-11\vec {\omega });\)

  • \({\mathcal O}(\vec {x}_2+\vec {x}_3)[-1]={\mathcal O}(-\vec {x}_1)={\mathcal O}(\vec {x}_2)(-7\vec {\omega })\).

Then the result follows by noting \([2]=(18\vec {\omega })\) in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\). \(\square \)

Moreover, by Proposition 5.4, we have

Lemma A.7

For any \(\vec {x}=\sum _{1\le i\le 3}l_i\vec {x}_i+l\vec {c}\) in normal form, the following hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\).

  1. (1)

    \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_2-\vec {x}), {\mathcal O}(\vec {x}_2))\ne 0\) if and only if \(0\le \vec {x}\le \vec {c}+\vec {x}_3\) and \(l_2=0\);

  2. (2)

    \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_2+\vec {x}_3-\vec {x}), {\mathcal O}(\vec {x}_2+\vec {x}_3))\ne 0\) if and only if \(0\le \vec {x}\le \vec {c}\) and \(l_1=0\).

Proposition A.8

\(T_{(2,4,7)}\) is extension-free in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\).

Proof

Recall that \(T_{(2,4,7)}=\big (\bigoplus \limits _{k=0}^{6}L(3k\vec {x}_3)\big )\oplus \big (\bigoplus \limits _{k=0}^{5}L((3k+1)\vec {x}_3)\big )\), where \(L={\mathcal O}(\vec {x}_2)\). For contradiction we assume \(T_{(2,4,7)}\) is not extension-free, then there exist some integers \(0\le a,b\le 18\) satisfying \(a,b\ne 3k+2\) for \(0\le k\le 5\), and \(0\ne m\in \mathbb {Z}\), such that

$$\begin{aligned} \underline{\textrm{Hom}}(L(a\vec {x}_{3}), L(b\vec {x}_{3})[m])\ne 0. \end{aligned}$$

Recall that \([2]=(18\vec {\omega })\) in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,4,7)\). We consider the following two cases.

Case 1: m is even, say, \(m=2n\). We get

$$\begin{aligned} \underline{\textrm{Hom}}(L(a\vec {x}_{3}), L(b\vec {x}_{3})[m])=\underline{\textrm{Hom}}(L(a\vec {x}_{3}-18n\vec {\omega }), L(b\vec {x}_{3})). \end{aligned}$$

For \(b=3k, 0\le k\le 6\), then according to Lemma A.7 (1), we have \(0\le (b-a)\vec {x}_{3}+18n\vec {\omega }\le \vec {c}+\vec {x}_{3}\) and the coefficient of \(\vec {x}_2\) in the normal form of \((b-a)\vec {x}_{3}+18n\vec {\omega }\) is zero. Hence 4|18n, i.e, n is even, say, \(n=2n'\). Then \((b-a)\vec {x}_{3}+18n\vec {\omega }=(b-a+27n')\vec {x}_3\). Thus we get \(0\le b-a+27n'\le 8\). Since \(0\le a, b\le 18\), we get \(n'=0\). It follows that \(n=0\) and then \(m=0\), a contradiction.

For \(b=3k+1, 0\le k\le 5\), then according to Lemma A.7 (2), we have \(0\le (b-a)\vec {x}_{3}+18n\vec {\omega }\le \vec {c}\). If n is even, say, \(n=2n'\), then \(0\le (b-a)\vec {x}_{3}+27n'\vec {x}_3 \le \vec {c}\), hence \(0\le b-a+27n'\le 7\). Since \(0\le a, b\le 18\), we get \(n'=0\). It follows that \(n=0\) and then \(m=0\), a contradiction. If n is odd, say, \(n=2n'-1\), then \(0\le (b-a)\vec {x}_{3}+18n\vec {\omega }=(b-a+27n'-17)\vec {x}_3+2\vec {x}_2\le \vec {c}\). It follows that \(b-a+27n'-17=0\). Hence \(a\equiv 2(\textrm{mod}\ 3)\), a contradiction.

Case 2: m is odd, say, \(m=2n+1\). By Serre duality we get

$$\begin{aligned} \underline{\textrm{Hom}}(L(a\vec {x}_{3}), L(b\vec {x}_{3})[2n+1])=D\underline{\textrm{Hom}}(L(b\vec {x}_{3}+18n\vec {\omega }-\vec {\omega }), L(a\vec {x}_{3})). \end{aligned}$$

According to Lemma A.7, by similar arguments as above we obtain that in the normal form of \((a-b)\vec {x}_3-18n\vec {\omega }+\vec {\omega }\), the coefficient of \(\vec {x}_1\) or \(\vec {x}_2\) is zero. It follows that \(2|18n-1\), a contradiction.

This finishes the proof. \(\square \)

1.3 A.3 The Weight Type (2,5,5)

Assume \(\mathbb {X}\) has weight type (2,5,5). Then the \(\delta \)-datum is given by \(\delta (\vec {c}; \vec {x}_1, \vec {x}_2, \vec {x}_3; \vec {\omega })=(10;5,2,2;1)\).

According to [24], the semigroup \(\{n\vec {\omega }| n\vec {\omega }\ge 0\}\) of \(\mathbb {Z}\vec {\omega }\) has minimal generating system \(\{4\vec {\omega }=\vec {x}_2+\vec {x}_3; 5\vec {\omega }=\vec {x}_1\}\), and the subalgebra \(R=S|_{\mathbb {Z}\vec {\omega }}\cong k[x,y,z]/(f)\), where \(x=x_2x_3, y=x_1, z=x_2^5\) and \(f=z^2+y^2z+x^5\). Moreover, R is \(\mathbb {Z}\)-graded in the sense that \(\deg (x,y,z;f)=(4,5,10;20)\). Thus the second suspension functor of \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\) is given by degree shift: \([2]=(20\vec {\omega })=(2\vec {c})\).

Since \(4\vec {\omega }=\vec {x}_2+\vec {x}_3>0\), the set \(\mathcal {S}\) defined in Eq. 5.1 has the form

$$\begin{aligned} \mathcal {S}=\{\vec {x}\, |\, 0\le \vec {x}\le n\vec {\omega }+\vec {c}, \text {for\ any} \ 2\le n\le 5\} =\{0, \vec {x}_2, \vec {x}_3, 2\vec {x}_2, 2\vec {x}_3\}. \end{aligned}$$

Hence each line bundle belongs to the orbit \(\tau ^{\mathbb {Z}}{\mathcal O}(j\vec {x}_i)\) for some \(0\le j\le 2\le i\le 3\).

By Proposition 5.3, the projective covers of \({\mathcal O}(\vec {x}_3)\) and \({\mathcal O}(2\vec {x}_3)\), under \(\tau ^{\mathbb {Z}}{\mathcal O}\)-exact structure on \(\textrm{vect}-\mathbb {X}\), are given by \(\mathcal {P}({\mathcal O}(\vec {x}_3))={\mathcal O}\oplus {\mathcal O}(-6\vec {\omega })\) and \(\mathcal {P}({\mathcal O}(2\vec {x}_3))={\mathcal O}\oplus {\mathcal O}(-2\vec {\omega })\), which fit into the following exact sequences:

(A.6)
(A.7)

Lemma A.9

The following statements hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\).

  1. (1)

    \({\mathcal O}(\vec {x}_3)[1]={\mathcal O}(\vec {x}_2)(10\vec {\omega });\)

  2. (2)

    \({\mathcal O}(2\vec {x}_3)[1]={\mathcal O}(2\vec {x}_2)(10\vec {\omega }).\)

Proof

Since we have exact sequences (A.6) and (A.7) for projective covers, the following hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\):

  • \({\mathcal O}(\vec {x}_3)[-1]={\mathcal O}(-4\vec {x}_2)={\mathcal O}(\vec {x}_2)(-10\vec {\omega });\)

  • \({\mathcal O}(2\vec {x}_3)[-1]={\mathcal O}(-3\vec {x}_2)={\mathcal O}(2\vec {x}_2)(-10\vec {\omega })\).

Then the result follows by noting \([2]=(20\vec {\omega })\) in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\). \(\square \)

Moreover, by Proposition 5.4, we have

Lemma A.10

The following statements hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\).

  1. (1)

    \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3-\vec {x}), {\mathcal O}(\vec {x}_3))\ne 0\) if and only if \(0\le \vec {x}\le \vec {x}_1+3\vec {x}_2\);

  2. (2)

    \(\underline{\textrm{Hom}}({\mathcal O}(2\vec {x}_3-\vec {x}), {\mathcal O}(2\vec {x}_3))\ne 0\) if and only if \(0\le \vec {x}\le \vec {x}_1+2\vec {x}_2+\vec {x}_3.\)

Proposition A.11

\(T_{(2,5,5)}\) is extension-free in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,5)\).

Proof

Recall that \(T_{(2,5,5)}=\bigoplus \limits _{a=1}^{4}\bigoplus \limits _{k=0}^{2}{\mathbb {S}}^k({\mathcal O}(a\vec {x}_3))\). For contradiction we assume \(T_{(2,5,5)}\) is not extension-free. Then by Lemma A.2 and using Serre duality, there exist some integers \(1\le a, b \le 4\), \(1\le k\le 3\) and \(0\ne m\in \mathbb {Z}\), such that

$$\begin{aligned} \underline{\textrm{Hom}}({\mathcal O}(a\vec {x}_{3}), {\mathbb {S}}^k({\mathcal O}(b\vec {x}_{3})[m]))\ne 0. \end{aligned}$$

Recall that \(\vec {c}=10\vec {\omega }\) and \([2]=(2\vec {c})=(20\vec {\omega })\). Hence, by Lemma A.9 we have

$$\begin{aligned} {\mathbb {S}}^k({\mathcal O}(b\vec {x}_{3})[m])= \left\{ \begin{array}{lll} {\mathcal O}(b\vec {x}_2+k\vec {\omega })((k+m)\vec {c}), &{}&{} \;\text {if}\; k+m \text {\;is\; odd};\\ {\mathcal O}(b\vec {x}_3+k\vec {\omega })((k+m)\vec {c}), &{}&{} \;\text {if}\; k+m \;\text {is\; even}. \end{array} \right. \end{aligned}$$

By Lemma A.10 (and its symmetric version by exchanging \(\vec {x}_2\) and \(\vec {x}_3\)), we have for \(i=2,3\),

$$\begin{aligned} 0\le b\vec {x}_i-a\vec {x}_3+k\vec {\omega }+(k+m)\vec {c}\le \left\{ \begin{array}{lll} \vec {x}_1+3\vec {x}_2, &{}&{} \text {\;if\;} {\mathcal O}(b\vec {x}_i)\in \tau ^\mathbb {Z}({\mathcal O}(\vec {x}_3));\\ \vec {x}_1+3\vec {x}_3, &{}&{} \text {\;if\;} {\mathcal O}(b\vec {x}_i)\in \tau ^\mathbb {Z}({\mathcal O}(\vec {x}_2));\\ \vec {x}_1+2\vec {x}_2+\vec {x}_3, &{}&{} \text {\;if\;} {\mathcal O}(b\vec {x}_i)\in \tau ^\mathbb {Z}({\mathcal O}(2\vec {x}_3));\\ \vec {x}_1+\vec {x}_2+2\vec {x}_3, &{}&{} \text {\;if\;} {\mathcal O}(b\vec {x}_i)\in \tau ^\mathbb {Z}({\mathcal O}(2\vec {x}_2)).\\ \end{array} \right. \end{aligned}$$
(A.8)

It follows that \(0\le 2(b-a)+k+10(k+m)\le 11\) by considering the degrees, which yields that \(k+m=0\) or 1.

Case 1: \(k+m=0\), then \(2\le k\le 3\) since \(m\ne 0\) and \((b-a)\vec {x}_3+k\vec {\omega }\ge 0\), which implies that \((b-a)\vec {x}_3\ge 2\vec {x}_3\), i.e., \(b-a\ge 2\). Hence \(b=3\) or 4. If \(b=3\), then \(a=1\). Observe that \(3\vec {x}_3=2\vec {x}_2+2\vec {\omega }\), i.e., \({\mathcal O}(3\vec {x}_3)\in \tau ^\mathbb {Z}({\mathcal O}(2\vec {x}_2))\). Thus \(2\vec {x}_3+k\vec {\omega }\le \vec {x}_1+\vec {x}_2+2\vec {x}_3\) by Eq. A.8, a contradiction to \(2\le k\le 3\). If \(b=4\), then \(a=1\) or 2. Observe that \(4\vec {x}_3=\vec {x}_2+6\vec {\omega }\). By Eq. A.8 we have \(0\le (4-a)\vec {x}_3+k\vec {\omega }\le \vec {x}_1+3\vec {x}_3\), contradicting to \(2\le k\le 3\).

Case 2: \(k+m=1\), then \(m\ne 0\) implies \(2\le k\le 3\). Moreover, \(0\le 2(b-a)+k+10\le 11\) implies that \(b<a\).

  • If \(b=1\), then by Eq. A.8 we have \(0\le \vec {x}_2-a\vec {x}_3+k\vec {\omega }+\vec {c}\le \vec {x}_1+3\vec {x}_3\). By considering the normal forms, we see that the coefficient of \(\vec {x}_2\) for the middle term equals zero, which implies \(k=1\), a contradiction.

  • If \(b=2\), then by Eq. A.8 we have \(0\le 2\vec {x}_2-a\vec {x}_3+k\vec {\omega }+\vec {c}\le \vec {x}_1+\vec {x}_2+2\vec {x}_3\). By considering the coefficient of \(\vec {x}_2\) in the normal form for the middle term, we get \(k=2\). In this case, \(2\vec {x}_2-a\vec {x}_3+2\vec {\omega }+\vec {c}=(8-a)\vec {x}_3\le \vec {x}_1+\vec {x}_2+2\vec {x}_3\), a contradiction.

  • If \(b=3\), then \(a=3\) or 4. Since \(3\vec {x}_2=2\vec {x}_3+2\vec {\omega }\), by Eq. A.8 we have \(0\le 3\vec {x}_2-a\vec {x}_3+k\vec {\omega }+\vec {c}\le \vec {x}_1+2\vec {x}_2+\vec {x}_3\). By considering the coefficient of \(\vec {x}_3\) in the normal form for the middle term, we get \(k=2\). In this case, \(3\vec {x}_2-a\vec {x}_3+2\vec {\omega }+\vec {c}=\vec {x}_2+(8-a)\vec {x}_3\le \vec {x}_1+2\vec {x}_2+\vec {x}_3\), a contradiction.

This finishes the proof. \(\square \)

1.4 A.4 The Weight Type (2,5,6)

Assume \(\mathbb {X}\) has weight type (2,5,6). Then the \(\delta \)-datum is given by \(\delta (\vec {c}; \vec {x}_1, \vec {x}_2, \vec {x}_3; \vec {\omega })=(30;15,6,5;4)\).

According to [24], \(\{n\vec {\omega }| n\vec {\omega }\ge 0\}\) of \(\mathbb {Z}\vec {\omega }\) has minimal generating system \(\{4\vec {\omega }=\vec {x}_2+2\vec {x}_3; 5\vec {\omega }=\vec {x}_1+\vec {x}_3; 6\vec {\omega }=4\vec {x}_2\}\), and the subalgebra \(R=S|_{\mathbb {Z}\vec {\omega }}\cong k[x,y,z]/(f)\), where \(x=x_2x_3^2, y=x_1x_3, z=x_2^4\) and \(f=xz^2+y^2z+x^4\). Moreover, R is \(\mathbb {Z}\)-graded in the sense that \(\deg (x,y,z;f)=(4,5,6;16)\). Thus the second suspension functor of \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\) is given by degree shift: \([2]=(16\vec {\omega })=(\vec {c}+4\vec {x}_2+2\vec {x}_3)\).

Since \(4\vec {\omega }=\vec {x}_2+2\vec {x}_3>0\), the set \(\mathcal {S}\) defined in Eq. 5.1 has the form

$$\begin{aligned} \mathcal {S}=\{\vec {x}\, |\, 0\le \vec {x}\le n\vec {\omega }+\vec {c}, \text {\ for\ any \ } 2\le n\le 5\} =\{\vec {x}\, |\, 0\le \vec {x}\le 3\vec {x}_3 \text {\ or } 0\le \vec {x}\le 2\vec {x}_2+\vec {x}_3 \}. \end{aligned}$$

Hence each line bundle belongs to the orbit \(\tau ^{\mathbb {Z}}{\mathcal O}(\vec {x})\) for some \(0\le \vec {x}\le 3\vec {x}_3\) or \(0\le \vec {x}\le 2\vec {x}_2+\vec {x}_3\).

By Proposition 5.3, the projective covers of \({\mathcal O}(j\vec {x}_i)\) for \(1\le j\le 2\le i\le 3\) are given by the middle terms of the following exact sequences respectively:

(A.9)
(A.10)
(A.11)
(A.12)

Lemma A.12

The following statements hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\).

  1. (1)

    \({\mathcal O}(\vec {x}_2)[1]={\mathcal O}(2\vec {x}_3)(6\vec {\omega });\)

  2. (2)

    \({\mathcal O}(\vec {x}_3)[1]={\mathcal O}(\vec {x}_2+\vec {x}_3)(5\vec {\omega });\)

  3. (3)

    \({\mathcal O}(2\vec {x}_2)[1]={\mathcal O}(2\vec {x}_2)(8\vec {\omega });\)

  4. (4)

    \({\mathcal O}(2\vec {x}_3)[1]={\mathcal O}(\vec {x}_2)(10\vec {\omega }).\)

Proof

Since we have exact sequences (A.9)–(A.12) for projective covers, the following hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\):

  • \({\mathcal O}(\vec {x}_2)[-1]={\mathcal O}(-\vec {c})={\mathcal O}(2\vec {x}_3)(-10\vec {\omega });\)

  • \({\mathcal O}(\vec {x}_3)[-1]={\mathcal O}(-\vec {x}_1-3\vec {x}_2)={\mathcal O}(\vec {x}_2+\vec {x}_3)(-11\vec {\omega });\)

  • \({\mathcal O}(2\vec {x}_2)[-1]={\mathcal O}(-4\vec {x}_3)={\mathcal O}(2\vec {x}_2)(-8\vec {\omega });\)

  • \({\mathcal O}(2\vec {x}_3)[-1]={\mathcal O}(-3\vec {x}_2)={\mathcal O}(\vec {x}_2)(-6\vec {\omega }).\)

Then the result follows by noting \([2]=(16\vec {\omega })\) in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\). \(\square \)

Moreover, by Proposition 5.4, we have

Lemma A.13

For any \(\vec {x}=\sum _{1\le i\le 3}l_i\vec {x}_i+l\vec {c}\) in normal form, the following hold in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\).

  1. (1)

    \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_2-\vec {x}), {\mathcal O}(\vec {x}_2))\ne 0\) if and only if \(0\le \vec {x}\le \vec {x}_1+5\vec {x}_3\);

  2. (2)

    \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3-\vec {x}), {\mathcal O}(\vec {x}_3))\ne 0\) if and only if \(0\le \vec {x}\le \vec {c}+2\vec {x}_2 \text {\ and\ } l_3=0\);

  3. (3)

    \(\underline{\textrm{Hom}}({\mathcal O}(2\vec {x}_2-\vec {x}), {\mathcal O}(2\vec {x}_2))\ne 0\) if and only if \(0\le \vec {x}\le \vec {x}_1+\vec {x}_2+3\vec {x}_3;\)

  4. (4)

    \(\underline{\textrm{Hom}}({\mathcal O}(2\vec {x}_3-\vec {x}), {\mathcal O}(2\vec {x}_3))\ne 0\) if and only if \(0\le \vec {x}\le \vec {x}_1+2\vec {x}_2+\vec {x}_3.\)

Proposition A.14

\(T_{(2,5,6)}\) is extension-free in \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\).

Proof

Recall that \(T_{(2,5,6)}=\big (\bigoplus \limits _{k=0}^{3}{\mathbb {S}}^k({\mathcal O}(\vec {x}_3))\big )\oplus \big ( \bigoplus \limits _{k=0}^{2}{\mathbb {S}}^k({\mathcal O}(2\vec {x}_3)\oplus {\mathcal O}(4\vec {x}_3)\oplus {\mathcal O}(6\vec {x}_3))\big ).\) In \(\underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\) we have \([2]=(16\vec {\omega })\), \({\mathcal O}(4\vec {x}_3)={\mathcal O}(2\vec {x}_2)(2\vec {\omega })\) and \({\mathcal O}(6\vec {x}_3)={\mathcal O}(\vec {x}_2)(6\vec {\omega })\). For any objects \(X,Y\in \underline{\textrm{vect}}^{\mathbb {Z}}-\mathbb {X}(2,5,6)\), by Serre duality we have

$$\begin{aligned} \underline{\textrm{Hom}}(X, Y[2n+1])=D\underline{\textrm{Hom}}(Y[2n], \tau X)=D\underline{\textrm{Hom}}(Y, X((1-16n)\vec {\omega })). \end{aligned}$$

Denote by \(V_0=\bigoplus \limits _{k=0}^{3}{\mathbb {S}}^{k}({\mathcal O}(\vec {x}_3))\) and \(V_i=\bigoplus \limits _{k=0}^{2}{\mathbb {S}}^{k}({\mathcal O}(2i\vec {x}_3))\) for \(1\le i\le 3\). Then \(T=V_0\oplus V_1\oplus V_2\oplus V_3\). We consider the following steps.

  1. (i)

    \(V_0\) is extension-free. For contradiction, by Lemma A.1 there exist \(m\ne 0\) and \(1\le k\le 4\), such that \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3), \tau ^{k}{\mathcal O}(\vec {x}_3)[k+m])\ne 0\). By Lemma A.13 we have

    • \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3), {\mathcal O}(\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0\) or 6.

    If \(k+m=2n\) for some n, then \(k+16n=0\) or 6, it follows that \(k=n=0\), a contradiction; if \(k+m=2n+1\) for some n, then \(1-(k+16n)=0\) or 6, it follows that \(n=0\) and \(k=1\), and then \(m=0\), a contradiction.

  2. (ii)

    \(V_i\) is extension-free for any \(1\le i\le 3\). For contradiction, by Lemma A.1 there exist \(m\ne 0\) and \(1\le k\le 3\), such that \(\underline{\textrm{Hom}}({\mathcal O}(2i\vec {x}_3), \tau ^{k}{\mathcal O}(2i\vec {x}_3)[k+m])\ne 0\). By Lemma A.13 we have

    • \(\underline{\textrm{Hom}}({\mathcal O}(2\vec {x}_3), {\mathcal O}(2\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0\) or 5;

    • \(\underline{\textrm{Hom}}({\mathcal O}(4\vec {x}_3), {\mathcal O}(4\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0,4,5\) or 9;

    • \(\underline{\textrm{Hom}}({\mathcal O}(6\vec {x}_3), {\mathcal O}(6\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0\) or 5.

    If \(k+m=2n\) for some n, then \(k+16n\in \{0,4,5,9\}\), it follows that \(k=n=0\), a contradiction; if \(k+m=2n+1\) for some n, then \(1-(k+16n)\in \{0,4,5,9\}\), it follows that \(n=0, k=1\), and then \(m=0\), a contradiction.

  3. (iii)

    \(V_0\oplus V_i\) is extension-free for any \(1\le i\le 3\). For contradiction, by Lemma A.2 there exist \(m\ne 0\) and \(-3\le k\le 3\), such that \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3), \tau ^{k}{\mathcal O}(2i\vec {x}_3)[k+m])\ne 0\). By Lemma A.13 we have

    • \(\underline{\textrm{Hom}}({\mathcal O}(\vec {x}_3), {\mathcal O}(2i\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0\);

    • \(\underline{\textrm{Hom}}({\mathcal O}(2i\vec {x}_3), {\mathcal O}(\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=4i+1\).

    If \(k+m=2n\) for some n, then \(k+16n=0\), it follows that \(k=n=0\), then \(m=0\), a contradiction; if \(k+m=2n+1\) for some n, then \(1-(k+16n)=4i+1\), it follows that \(n=0\), then \(k=-4i\), a contradiction.

  4. (iv)

    \(V_i\oplus V_j\) is extension-free for any \(1\le i<j\le 3\). For contradiction, by Lemma A.2 there exist \(m\ne 0\) and \(-2\le k\le 3\), such that \(\underline{\textrm{Hom}}({\mathcal O}(2i\vec {x}_3), \tau ^{k}{\mathcal O}(2j\vec {x}_3)[k+m])\ne 0\). By Lemma A.13 we have

    • \(\underline{\textrm{Hom}}({\mathcal O}(2i\vec {x}_3), {\mathcal O}(2j\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=0\) or 5;

    • \(\underline{\textrm{Hom}}({\mathcal O}(2j\vec {x}_3), {\mathcal O}(2i\vec {x}_3+n\vec {\omega }))\ne 0\) if and only if \(n=4(j-i)\) or \(4(j-i)+5\).

    If \(k+m=2n\) for some n, then \(k+16n=0\) or 5, it follows that \(k=n=0\), then \(m=0\), a contradiction; if \(k+m=2n+1\) for some n, then \(1-(k+16n)=4(j-i)\) or \(4(j-i)+5\), it follows that \(n=0\) and \(k=1-4(j-i)\) or \(-4-4(j-i)\), a contradiction.

Then the proof is finished. \(\square \)

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

Lenzing, H., Meltzer, H. & Ruan, S. Nakayama Algebras and Fuchsian Singularities. Algebr Represent Theor 27, 815–846 (2024). https://doi.org/10.1007/s10468-023-10236-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10468-023-10236-8

Keywords

Mathematics Subject Classification (2010)

Navigation