Skip to main content
Log in

Noncommutative products of Euclidean spaces

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

Abstract

We present natural families of coordinate algebras on noncommutative products of Euclidean spaces \({\mathbb {R}}^{N_1} \times _{\mathcal {R}} {\mathbb {R}}^{N_2}\). These coordinate algebras are quadratic ones associated with an \(\mathcal {R}\)-matrix which is involutive and satisfies the Yang–Baxter equations. As a consequence, they enjoy a list of nice properties, being regular of finite global dimension. Notably, we have eight-dimensional noncommutative euclidean spaces \({\mathbb {R}}^{4} \times _{\mathcal {R}} {\mathbb {R}}^{4}\). Among these, particularly well behaved ones have deformation parameter \(\mathbf{u} \in {\mathbb {S}}^2\). Quotients include seven spheres \({\mathbb {S}}^{7}_\mathbf{u}\) as well as noncommutative quaternionic tori \({\mathbb {T}}^{{\mathbb {H}}}_\mathbf{u} = {\mathbb {S}}^3 \times _\mathbf{u} {\mathbb {S}}^3\). There is invariance for an action of \({{\mathrm{SU}}}(2) \times {{\mathrm{SU}}}(2)\) on the torus \({\mathbb {T}}^{{\mathbb {H}}}_\mathbf{u}\) in parallel with the action of \(\mathrm{U}(1) \times \mathrm{U}(1)\) on a ‘complex’ noncommutative torus \({\mathbb {T}}^2_\theta \) which allows one to construct quaternionic toric noncommutative manifolds. Additional classes of solutions are disjoint from the classical case.

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. Berger, R.: Confluence and quantum Yang-Baxter equation. J. Pure Appl. Algebra 145(2000), 267–283 (2000)

    Article  MathSciNet  Google Scholar 

  2. Berger, R.: Dimension de Hochschild des algèbres gradués. C. R. Math. Acad. Sci. Paris 341, 597–600 (2005)

    Article  MathSciNet  Google Scholar 

  3. Berger, R., Ginzburg, V.: Higher symplectic reflection algebras and non-homogeneous \({N}\)-Koszul property. J. Algebra 305, 577–601 (2006)

    Article  MathSciNet  Google Scholar 

  4. Braverman, A., Gaitsgory, D.: Poincaré–Birkhoff–Witt theorem for quadratic algebras of Koszul type. J. Algebra 181, 315–328 (1996)

    Article  MathSciNet  Google Scholar 

  5. Cartan, H.: Homologie et cohomologie d’une algèbre graduée. Séminaire Henri Cartan 11(2), 1–20 (1958)

    Google Scholar 

  6. Connes, A., Dubois-Violette, M.: Noncommutative finite-dimensional manifolds. I. Spherical manifolds and related examples. Commun. Math. Phys. 230, 539–579 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  7. Connes, A., Landi, G.: Noncommutative manifolds, the instanton algebra and isospectral deformations. Commun. Math. Phys. 221, 141–159 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  8. Dubois-Violette, M.: Multilinear forms and graded algebras. J. Algebra 317, 198–225 (2007)

    Article  MathSciNet  Google Scholar 

  9. Dubois-Violette, M.: Poincaré duality for Koszul algebras. Algebra, Geometry and Mathematical Physics. Springer Proceedings in Mathematics and Statistics, vol. 85, pp. 3–26. Springer, Berlin (2014)

    Google Scholar 

  10. Dubois-Violette, M., Landi, G.: Noncommutative Euclidean spaces. arXiv:1801.03410 [math.QA]

  11. Ginzburg, V.: Calabi–Yau Algebras. arXiv:math/0612139 [math.AG]

  12. Gurevich, D.I.: Algebraic aspects of the quantum Yang–Baxter equation. Algebra i Analiz 2, 119–148 (1990)

    MATH  Google Scholar 

  13. Manin, Y.I.: Quantum Groups and Non-commutative Geometry. CRM Université de Montréal, Montreal (1988)

    MATH  Google Scholar 

  14. Polishchuk, A., Positselski, L.: Quadratic algebras. University Lecture Series, vol. 37. American Mathematical Society, Providence (2005)

    MATH  Google Scholar 

  15. Positselski, L.: Nonhomogeneous quadratic duality and curvature. Funct. Anal. Appl. 27, 197–204 (1993)

    Article  MathSciNet  Google Scholar 

  16. Wambst, M.: Complexes de Koszul quantiques. Ann. Inst. Fourier Grenoble 43, 1089–1156 (1993)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giovanni Landi.

Appendix A: Quadratic algebras

Appendix A: Quadratic algebras

To be definite, we take the ground field to be complex numbers \({\mathbb {C}}\). A homogeneous quadratic algebra [13, 14] is an associative algebra \(\mathcal {A}\) of the form

$$\begin{aligned} \mathcal {A}=A(E,R)=T(E)/(R) \end{aligned}$$

with E a finite-dimensional vector space, R a subspace of \(E\otimes E\) and (R) denoting the two-sided ideal of the tensor algebra T(E) over E generated by R. The space E is the space of generators of \(\mathcal {A}\) and the subspace R of \(E\otimes E\) is the space of relations of \(\mathcal {A}\). The algebra \(\mathcal {A}=A(E,R)\) is naturally a graded algebra \(\mathcal {A}=\bigoplus _{n\in \mathbb N}\mathcal {A}_n\) which is connected, that is such that \(\mathcal {A}_0={\mathbb {C}}{\mathbb {1}}\) and generated by the degree 1 part, \(\mathcal {A}_1=E\).

To a quadratic algebra \(\mathcal {A}=A(E,R)\) as above one associates another quadratic algebra, its Koszul dual \(\mathcal {A}^!\), defined by

$$\begin{aligned} \mathcal {A}^!=A\left( E^*,R^\perp \right) \end{aligned}$$

where \(E^*\) denotes the dual vector space of E and \(R^\perp \subset E^*\otimes E^*\) is the orthogonal of the space of relations \(R\subset E\otimes E\) defined by

$$\begin{aligned} R^\perp = \left\{ \omega \in E^*\otimes E^*\, ; \, \langle \omega ,r\rangle =0,\forall r\in R\right\} . \end{aligned}$$

As usual, by using the finite-dimensionality of E, one identifies \(E^*\otimes E^*\) with the dual vector space \((E\otimes E)^*\) of \(E\otimes E\). One has of course \((\mathcal {A}^!)^!=\mathcal {A}\) and the dual vector spaces \(\mathcal {A}^{!*}_n\) of the homogeneous components \(\mathcal {A}^!_n\) of \(\mathcal {A}^!\) are

$$\begin{aligned} \mathcal {A}^{!*}_1=E \quad \hbox {and} \quad \mathcal {A}^{!*}_n=\bigcap _{r+s+2=n} E^{\otimes ^r}\otimes R \otimes E^{\otimes ^s} \end{aligned}$$
(A.1)

for \(n\ge 2\), as easily verified. In particular \(\mathcal {A}^{!*}_2=R\) and \(\mathcal {A}^{!*}_n\subset E^{\otimes ^n}\) for any \(n\in \mathbb N\).

Consider the sequence of free left \(\mathcal {A}\)-modules

$$\begin{aligned} K(\mathcal {A}): \quad \cdots {\mathop {\rightarrow }\limits ^{b}}\mathcal {A}\otimes \mathcal {A}^{!*}_{n+1}{\mathop {\rightarrow }\limits ^{b}}\mathcal {A}\otimes \mathcal {A}^{!*}_n\rightarrow \cdots \rightarrow \mathcal {A}\otimes \mathcal {A}^{!*}_2{\mathop {\rightarrow }\limits ^{b}}\mathcal {A}\otimes E{\mathop {\rightarrow }\limits ^{b}}\mathcal {A}\rightarrow 0 \end{aligned}$$
(A.2)

where \(b:\mathcal {A}\otimes \mathcal {A}^{!*}_{n+1}\rightarrow \mathcal {A}\otimes \mathcal {A}^{!*}_n\) is induced by the left \(\mathcal {A}\)-module homomorphism of \(\mathcal {A}\otimes E^{\otimes ^{n+1}}\) into \(\mathcal {A}\otimes E^{\otimes ^n}\) defined by

$$\begin{aligned} b(a\otimes (x_0\otimes x_1\otimes \cdots \otimes x_n))=ax_0 \otimes (x_1\otimes \cdots \otimes x_n) \end{aligned}$$

for \(a\in \mathcal {A}\), \(x_k \in E\). It follows from (A.1) that \(\mathcal {A}^{!*}_n\subset R\otimes E^{\otimes ^{n-2}}\) for \(n\ge 2\), which implies that \(b^2=0\). As a consequence, the sequence \(K(\mathcal {A})\) in (A.2) is a chain complex of free left \(\mathcal {A}\)-modules called the Koszul complex of the quadratic algebra \(\mathcal {A}\). The quadratic algebra \(\mathcal {A}\) is said to be a Koszul algebra whenever its Koszul complex is acyclic in positive degrees, that is, whenever \(H_n(K(\mathcal {A}))=0\) for \(n\ge 1\). One shows easily that \(\mathcal {A}\) is a Koszul algebra if and only if its Koszul dual \(\mathcal {A}^!\) is a Koszul algebra.

It is important to realize that the presentation of \(\mathcal {A}\) by generators and relations is equivalent to the exactness of the sequence

$$\begin{aligned} \mathcal {A}\otimes R {\mathop {\rightarrow }\limits ^{b}} \mathcal {A}\otimes E {\mathop {\rightarrow }\limits ^{b}}\mathcal {A}{\mathop {\rightarrow }\limits ^{\varepsilon }} {\mathbb {C}}\rightarrow 0 \end{aligned}$$

with \(\varepsilon \) the map induced by the projection onto degree 0. Thus one always has

$$\begin{aligned} H_1(K(\mathcal {A}))=0 \quad \hbox {and} \quad H_0(K(\mathcal {A}))={\mathbb {C}}\end{aligned}$$

and, whenever \(\mathcal {A}\) is Koszul, the sequence

$$\begin{aligned} K(\mathcal {A}){\mathop {\rightarrow }\limits ^{\varepsilon }} {\mathbb {C}}\rightarrow 0 , \end{aligned}$$

is a free resolution of the trivial module \({\mathbb {C}}\). This resolution is then a minimal projective resolution of \({\mathbb {C}}\) in the category of graded modules [5].

Let \(\mathcal {A}=A(E,R)\) be a quadratic Koszul algebra such that \(\mathcal {A}^!_D\not =0\) and \(\mathcal {A}^!_n=0\) for \(n>D\). Then the trivial (left) module \({\mathbb {C}}\) has projective dimension D which implies that \(\mathcal {A}\) has global dimension D (see [5]). This also implies that the Hochschild dimension of \(\mathcal {A}\) is D (see [2]). By applying the functor \({{\mathrm{Hom}}}_\mathcal {A}(\, \cdot \, , \mathcal {A})\) to the Koszul chain complex \(K(\mathcal {A})\) of left \(\mathcal {A}\)-modules one obtains the cochain complex \(L(\mathcal {A})\) of right \(\mathcal {A}\)-modules

$$\begin{aligned} L(\mathcal {A}) : \quad \quad 0\rightarrow \mathcal {A}{\mathop {\rightarrow }\limits ^{b'}} \cdots {\mathop {\rightarrow }\limits ^{b'}}\mathcal {A}^!_n \otimes \mathcal {A}{\mathop {\rightarrow }\limits ^{b'}} \mathcal {A}^!_{n+1} \otimes \mathcal {A}{\mathop {\rightarrow }\limits ^{b'}} \cdots . \end{aligned}$$

Here \(b'\) is the left multiplication by \(\sum _k\theta ^k\otimes e_k\) in \(\mathcal {A}^!\otimes \mathcal {A}\) where \((e_k)\) is a basis of E with dual basis \((\theta ^k)\). The algebra \(\mathcal {A}\) is said to be Koszul–Gorenstein if it is Koszul of finite global dimension D as above and if \(H^n(L(\mathcal {A}))={\mathbb {C}}\, \delta ^n_D\). Notice that this implies that \(\mathcal {A}^!_n\simeq \mathcal {A}^{!*}_{D-n}\) as vector spaces (this is a version of Poincaré duality).

Finally, a graded algebra \(\mathcal {A}=\oplus _n \mathcal {A}_n\) is said to have polynomial growth whenever there are a positive C and \(N \in {\mathbb {N}}\) such that, for any \(n \in {\mathbb {N}}\),

$$\begin{aligned} \dim (\mathcal {A}_n)\le C n^{N-1}. \end{aligned}$$

As before, let E be a finite-dimensional vector space with the tensor algebra T(E) endowed with its natural filtration \(F^n(T(E))=\bigoplus _{m\le n} E^{\otimes ^m}\). A nonhomogeneous quadratic algebra [4, 15], is an algebra \({\mathfrak {A}}\) of the form

$$\begin{aligned} {\mathfrak {A}}=A(E,P)=T(E)/(P) \end{aligned}$$

where P is a subspace of \(F^2(T(E))\) and where (P) denotes as above the two-sided ideal of T(E) generated by P. The filtration of the tensor algebra T(E) induces a filtration \(F^n({\mathfrak {A}})\) of \({\mathfrak {A}}\) and one associates to \({\mathfrak {A}}\) the graded algebra

$$\begin{aligned} \hbox {gr} ({\mathfrak {A}})=\oplus _n F^n({\mathfrak {A}})/F^{n-1}({\mathfrak {A}}). \end{aligned}$$

Let R be the image of P under the canonical projection of \(F^2(T(E))\) onto \(E\otimes E\) and let \(\mathcal {A}=A(E,R)\) be the homogeneous quadratic algebra T(E) / (R); this \(\mathcal {A}\) is referred to as the quadratic part of \({\mathfrak {A}}\). There is a canonical surjective-graded algebra homomorphism

$$\begin{aligned} \hbox {can} :\mathcal {A}\rightarrow \hbox {gr}({\mathfrak {A}}). \end{aligned}$$

One says that \({\mathfrak {A}}\) has the Poincaré–Birkhoff–Witt (PBW) property whenever this homomorphism is an isomorphism. The terminology comes from the example where \({\mathfrak {A}}=U({\mathfrak {g}})\) the universal enveloping algebra of a Lie algebra \({\mathfrak {g}}\). A central result (see [4, 14]) states that if \({\mathfrak {A}}\) has the PBW property then the following conditions are satisfied:

$$\begin{aligned}&\mathrm {(i)} \quad P\cap F^1 (T(E))=0 , \nonumber \\&\mathrm {(ii)} \quad (P \cdot E + E \cdot P)\cap F^2 (T(E)) \subset P. \end{aligned}$$
(A.3)

Conversely, if the quadratic part \(\mathcal {A}\) is a Koszul algebra, the conditions \(\mathrm {(i)}\) and \(\mathrm {(ii)}\) imply that \({\mathfrak {A}}\) has the PBW property. Condition (i) means that P is obtained from R by adding to each nonzero element of R terms of degrees 1 and 0. That is there are linear mappings \(\psi _1:R\rightarrow E\) and \(\psi _0:R \rightarrow {\mathbb {C}}\) such that one has

$$\begin{aligned} P=\{r+\psi _1(r)+\psi _0(r) {\mathbb {1}}\,\, \vert \,\, r \in R\} \end{aligned}$$
(A.4)

giving P in terms of R. Condition (ii) is a generalization of the Jacobi identity (see [14]).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dubois-Violette, M., Landi, G. Noncommutative products of Euclidean spaces. Lett Math Phys 108, 2491–2513 (2018). https://doi.org/10.1007/s11005-018-1090-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-018-1090-z

Keywords

Mathematics Subject Classification

Navigation