The geometry of sedenion zero divisors
Abstract.
The sedenion algebra is a non-commutative, non-associative, -dimensional real algebra with zero divisors. It is obtained from the octonions through the Cayley-Dickson construction. The zero divisors of can be viewed as the submanifold of normalized pairs whose product equals zero, or as the submanifold of normalized elements with non-trivial annihilators. We prove that is isometric to the excepcional Lie group , equipped with a naturally reductive left-invariant metric. Moreover, is the total space of a Riemannian submersion over the excepcional symmetric space of quaternion subalgebras of the octonion algebra, with fibers that are locally isometric to a product of two round -spheres with different radii. Additionally, we prove that is isometric to the Stiefel manifold , the space of orthonormal -frames in , endowed with a specific -invariant metric. By shrinking this metric along a circle fibration, we construct new examples of an Einstein metric and a family of homogenous metrics on with non-negative sectional curvature.
Key words and phrases:
Cayley-Dickson algebras, Sedenion algebra, Zero divisors, Einstein manifolds, Non-negative curvature2020 Mathematics Subject Classification:
53C30, 17A201. Introduction
The Cayley-Dickson algebras form a sequence of real algebras , defined recursively beginning with and doubling in dimension with each iteration. The first members of this family are the familiar real division algebras: , , and . The next algebra in this sequence is the so-called sedenion algebra , which is often overlooked in comparison to its lower-dimensional relatives due to its lack of certain desirable algebraic properties. Nonetheless, this somewhat enigmatic algebra has long intrigued mathematicians and has recently found applications in fields such as theoretical physics [GG19] and machine learning [SA20].
Since is not a division algebra, it is interesting to understand the structure of its zero divisors. The topology of the sedenion zero divisors is described by the principal bundle
where is the excepcional compact Lie group of rank and is the Stiefel manifold of orthonormal -frames in . Specifically, is homeomorphic to the submanifold of normalized sedenion pairs that multiply to zero; is homeomorphic to the submanifold of sedenions with norm that have non-trivial annihilators; and, for each , the fiber corresponds to the sphere of the annihilator subspace of (see [Mor98, BDI08]). However, little is known regarding the geometry of the sedenion zero divisors.
Both and carry a natural geometry as submanifolds of and , respectively. Furthermore, since the zero divisors of are invariant under , whose connected component is isomorphic to , it follows that and are homogeneous submanifolds. In this article, we study the intrinsic geometry of the zero divisors of . First, we prove that is isometric to with a naturally reductive left-invariant metric, forming the total space of a Riemannian submersion over the exceptional symmetric space , with fibers locally isometric to a product of two round -spheres with different radii.
Next, we analyze the geometry of , which is isometric to with a particular -invariant metric. At first glance, the geometry of does not seem very interesting; however, by shrinking the metric along a certain circle fibration, we obtain a family () of -invariant metrics on , where represents the original metric. This process reveals several distinguished examples. Specifically, we prove, among other things, that :
-
β’
has positive scalar curvature if and only of ;
-
β’
is an Einstein manifold if and only if ;
-
β’
has non-negative sectional curvature if and only if .
These results are quite remarkable, as Einstein metrics and metrics with non-negative curvature are very rare. To the best of our knowledge, the examples presented in this article are new. The known homogeneous Einstein metrics on are limited to the unique -invariant Einstein metric discovered by Sagle and the so-called Jensen metrics (see [Sag70, Jen73, BH87, Ker98]). It is worth noticing that the metric is neither -invariant nor a Jensen metric. Regarding metrics with non-negative sectional curvature, we refer to the survey [Zil07]. Typically, examples of homogeneous metrics with non-negative curvature appear as normal homogeneous metrics or are constructed through a Cheeger deformation of a metric already known to have non-negative curvature. Recall that none of the metrics is normal homogeneous (nor even naturally reductive) and that the initial metric does not possess non-negative sectional curvature.
Let us comment briefly on the proof our main results. In order to study the geometry of , it is necessary to βfix an originβ so that the metric can be identified with a left-invariant metric on . Any choice of such an origin for leads to isometric metrics on , but a well-chosen origin can greatly simplify computations. We select the origin from among the so-called 84 standard zero divisors of . Then, using the results in [DZ79], we show that is a naturally reductive space. A similar approach applies to the study of with the metric . Here, we select another standard zero divisor (different from the previous one) so that the isotropy subgroup of acts trivially on the usual subalgebra . This choice allows the metric to be expressed in diagonal form with respect to the normal homogeneous metric, making it possible to derive a nice expression for the Ricci tensor of .
The most challenging part is to determine the sign of the sectional curvatures of . Since there is no manageable expression for the curvature (as there is for naturally reductive spaces), using algebraic manipulation proves to be nearly impossible. Indeed, the sectional curvature function of can be interpreted as a homogeneous polynomial of degree in real variables, which, for a generic , has 285 non-trivial coefficients. To show that is non-negative for , we reduce the problem to proving that both (the formal extension of at ) and are non-negative. By using convex optimization techniques, we are able to prove the stronger result that and are polynomial sums of squares.
Finally we want to mention that the computations required in the proof of some of our results are often cumbersome and were computer checked using the software SageMath. The code used to verify our results is available at [Reg24].
We believe this work shows that the study of the geometry of Cayley-Dickson algebras, particularly regarding their zero divisors, deserves further attention, as it may have interesting implications in differential geometry of compact homogeneous spaces.
Acknowledgements
This work is supported by CONICET and partially supported by SeCyT-UNR and ANPCyT. The author would like to thank Andreas Arvanitoyeorgos for helpful discussions on homogeneous Einstein metrics on Stiefel manifolds.
2. Preliminaries and notation
The main references for this section are [Mor98, BDI08] on Cayley-Dickson algebras and their zero divisors, [Arv03] on the geometry of homogeneous spaces and [DZ79] on naturally reductive left-invariant metrics on compact Lie groups. Observe that in this section, as well as throughout the rest of the article, we start counting indices from 0.
2.1. Cayley-Dickson algebras
The Cayley-Dickson algebras are a family of real algebras, equipped with an involution (also called conjugation), which are recursively defined starting from , where . Each subsequent algebra is defined by setting as a vector space, with multiplication given by
and involution defined by
Notice that the inclusion is a monomorphism of algebras from into for all . It is well known that the first four algebras in the Cayley-Dickson construction are the real division algebras , , and , respectively. It is also known that the Cayley-Dickson algebras lose some important properties with each iteration. For example, is commutative if and only if , associative if and only if ; alterative (i.e., and for all ) if and only of . On the other hand, every Cayley-Dickson algebra is flexible (i.e., for all ) and power associative (i.e., is well defined for all and ).
For we define its real and imaginary parts as and , respectively. We say that is real (resp.Β imaginary) if (resp.Β ). Thus, one can recover the usual inner product on by
In , one has that for all . However, the identity does not hold in general if . Recall that is a division algebra if and only . If , then has zero divisors. Since, implies , the left and right zero divisors of coincide. Thus, an element is a zero divisor if and only if , where is the kernel of the -linear map given by . In [Mor98], it is proven that a zero divisor must be imaginary and . Furthermore, in [BDI08] it is proven that . One can study the zero divisors globally by defining the sets
Normalizing the zero divisors to is not particularly important, but it will be convenient later. When , the sets are also of interest.
For , one has that the automorphism group of is given by
where is the symmetric group in three elements, and is the -dimensional compact simple Lie group of rank . Recall that acts diagonally on . It follows that for all , where is the Lie algebra of .
2.2. Sedenion zero divisors
From now on we denote the sedenion algebra by . Let us denote by the canonical basis of . By making an abuse of notation, we also denote by and the canonical basis of and respectively. The zero divisors of have the following form.
Proposition 2.1 (See [BDI08]).
An element is a zero divisor if and only if are imaginary elements of such that and .
From this result, one can construct the standard zero divisors of . Namely, the elements of the form such that , , and (see Table 1). Clearly, every automorphism of maps into itself. Moreover, we have that the connected component of acts simply and transitively on :
Theorem 2.2 ([Mor98]).
is homeomorphic (and moreover, diffeomorphic) to .
Given , we have that . It is not difficult to see that the isotropy subgroup at is isomorphic to . Note that is diffeomorphic to the Stiefel manifold . In fact, every automorphism of is completely determined by its values at respectively. Here can be any triple of pairwise orthonormal imaginary octonions of norm 1 such that . Hence the map identifies with a transitive action of in , whose isotropy subgroup at are the octonion automorphism that act trivially on , and therefore are isomorphic to . Thus, the topology of the sedenion zero divisors is encoded by the principal bundle
2.3. The Lie algebra of
We think of the Lie group as a subgroup of in the natural way (since every automorphism of fixes , we have that is actually a subgroup of , but we do not use this identification here). So, we have as a subalgebra of . Let us consider the bi-invariant metric induced by the inner product on , which we denote with the same symbol, given by
Let us denote by , where , the matrix such that and in any other case. We define
One can see that is an orthonormal basis of with respect to the bi-invariant metric. We will denote by its dual basis. Define
We have that and are two subalgebras of isomorphic to such that . Moreover, is the subalgebra of a maximal subgroup of isomorphic to (cfr.Β [BLS20]). Such subgroup preserves the orthogonal decomposition . Furthermore, the subgroup of with Lie algebra is isomorphic to and acts trivially on . Recall that , with the normal homogeneous metric, is the symmetric space of quaternion subalgebras of .
2.4. Homogeneous and naturally reductive spaces
Let be a Lie group and be a compact subgroup of . Let us denote by and the Lie algebras of and , respectively. Assume that acts almost effectively on and that is endowed with a -invariant metric . Recall that every induces a Killing vector field on defined as . The map from into satisfies
Let us fix a reductive decomposition (i.e., is an -invariant subspace of complementary to ), which always exists since is compact. Assume that is the isotropy subgroup of . Then we can identify . The geometry of is determined by an -invariant inner product on , which we also denote by , defined such that the map is a linear isometry. With this setting, we can compute the Levi-Civita connection of as
(2.1) |
where is the algebraic tensor on given by
Let be the curvature tensor of . The sectional curvature of is determined by
for . Also, the Ricci tensor of is determined by
(2.2) |
for , where is an orthonormal basis of and .
Recall that the metric on is naturally reductive if and only if . An interesting particular case is when a left-invariant metric on a Lie group is naturally reductive (with respect to a certain transitive Lie group of isometries).
Theorem 2.3 ([DZ79]).
Let be a compact, simple Lie group group endowed with a left-invariant metric . Let denote the Lie algebra of and let be a bi-invariant metric on (which is a negative multiple of the Killing form of ). The metric is naturally reductive if and only if there exists a subalgebra of such that
where , with the center of and are simple ideals. Here, is the orthogonal complement of with respect to the bi-invariant metric, is an arbitrary inner product on , and are positive real numbers.
3. The -invariant metrics on and
Consider as a submanifold of with the induced metric. Although this reduction is not necessary here, one could lower the codimension of . In fact, by PropositionΒ 2.1, is a submanifold of . Since acts isometrically on , we have that is a homogeneous submanifold. Furthermore, by TheoremΒ 2.2, the diffeomorphism induces a left-invariant metric on .
Theorem 3.1.
The metric on is naturally reductive. Furthermore, is the total space of a Riemannian submersion over the excepcional symmetric space with totally geodesic fibers, which are locally isometric to a product of two round -spheres with different radii.
Proof.
It is sufficient to prove the theorem for the left-invariant metric on defined in the paragraph preceding the statement. To determine such a metric, one fixes an element and computes
(3.1) |
Note that not every zero divisor pair behaves nicely with respect to the decomposition given in Subsection 2.3. By running (3.1) over the standard zero divisors from Table 1, we observe that if the metric can be expressed as
From Theorem 2.3, it follows that this metric is naturally reductive. More precisely, this metric is naturally reductive with respect to , where the second factor acts on the right and the isotropy subgroup is given by . Thus, from [DZ79, Theorem 8], the subgroup , whose Lie algebra is given by , is totally geodesic.
Since is a multiple of the bi-invariant metric, when restricted to , and is orthogonal to with respect to both metrics, we conclude that is a Riemannian submersion. The fiber of this submersion is isometric to the Lie group endowed with the bi-invariant metric , which is obtained by taking two different scalings of the bi-invariant metric on the simple ideals of . Hence, the universal cover of splits into a product of two round spheres with different radii. β
Remark 3.2.
Since the metric in is naturally reductive, many geometric properties follow from existing results. For example, the (connected component of the) full isometry group is computed in [DZ79] (see also [OR13]). The so-called index of symmetry of , which in this case is trivial, can be computed from the results in [ORT14]. It can also be seen from [DZ79] that the metric on is not Einstein. We verify this fact again in the next proposition by explicitly computing the Ricci tensor, which also allows us to show that the Ricci curvature is positive.
Proposition 3.3.
has positive Ricci curvature. Moreover,
(3.2) |
Proof.
It follows from a straightforward computation using the following well-known formula. Let be an -orthonormal basis of . Then
where . Since has diagonal form in the basis , we can choose . From this, we can show that
which is equivalent to (3.2). β
Now we direct our attention to the geometry of with the metric induced from the ambient space . Since acts isometrically and transitively on , we have that is isometric to the Stiefel manifold , equipped with a certain -invariant metric, where is the isotropy subgroup of . We again denote by such a metric, which is defined by
Similarly to the case of , we can choose appropriately so that the Lie algebra of is . Taking , we obtain that
is a reductive decomposition for . The corresponding -invariant inner product on is given by
A detailed study of the isometry group and the curvature of is given in the next section. Before proceeding, we note a simple fact about the sectional curvatures of .
Remark 3.4.
Let us denote by the -dimensional subspace of generated by and , where . Then the sectional curvature of is non-negative if and only if . This suggests that one could attempt to modify the metric along the direction normal to inside in order to get some examples of metrics with non-negative sectional curvature. We explore this approach in the next section.
4. A family of -invariant metrics on
For each , we consider on the family of -invariant metrics given by
(4.1) |
Indeed, gives an -invariant inner product on since is the subspace of fixed points of the isotropy representation of and
Next, we compute the connected component of the full isometry group of .
Theorem 4.1.
.
Proof.
Since is a compact simple Lie group, it follows from the results in [Oni92] that , where is the normal homogeneous metric associated with the homogeneous presentation . From [Reg10], we have that (almost direct product) where the Lie algebra of is given by the -invariant vector fields, which are identified with the fixed vectors of the isotropy representation. That is, the Lie algebra of is identified with , but the elements of act βon the rightβ. Then, it is not difficult to see that (almost direct product) for a compact and connected subgroup of , which in principle depends on . Since , it is enough to see that and .
Now, for , let be the -invariant vector field induced by . Using (2.1) and the fact that for all , one can see that is a Killing field for if and only if is skew-symmetric with respect to . The - and -scalings of the metric on the irreducible subspaces and prevent and from being Killing fields for . However, one can check that is a Killing field for for any . Thus , which implies is actually a direct product. Observe that we have proved that the factor is independent of . β
Now, we compute the Ricci and scalar curvature of . In particular, we obtain the following result.
Theorem 4.2.
-
(1)
The metric is Einstein if and only if .
-
(2)
The metric has positive scalar curvature if and only if .
Proof.
Let be the -orthonormal basis of obtained from normalizing the basis . We can use formula (2.2) to explicitly compute the Ricci tensor of . After lengthy computations, carefully verified using a computer (see [Reg24]), we obtain
(4.2) |
Hence, is Einstein if and only if . Also, from (4.2) we get that the scalar curvature is positive if and only if . β
Remark 4.3.
In [Jen73], the construction of remarkable examples of Einstein metrics on the base space of certain principal bundles can be found. Such metrics are now known as Jensen metrics. In particular, there exist -invariant Einstein metrics on , arising from the principal bundle , which in our notation takes the form for certain values of . Notice that the metric from TheoremΒ 4.2 is not a Jensen metric. Moreover, it is not even bi-invariant when restricted to .
Theorem 4.4.
The metric has non-negative sectional curvature if and only if .
In order to prove our theorem, we will need the following result, which is a particular case of Theorem 1 in [PW98] (see also [CLR95]).
Lemma 4.5.
Let be a homogeneous polynomial of degree . Then is a (polynomial) sum of squares if and only if there exists a symmetric positive semi-definite matrix such that
(4.3) |
where is the vector of monomials of degree .
Let us mention that the vector has coordinates and the subspace of, not necessarily positive semi-definite, matrices satisfying (4.3) has dimension . Thus, finding an exact (positive semi-definite) solution for equation (4.3) can be quite difficult, even for relatively small values of .
Proof of Theorem 4.4.
It is not hard to see that if , then the sectional curvature of the plane is
Thus, does not have non-negative sectional curvature for . Let be the orthonormal basis of defined in the proof of Theorem 4.2 and write
For each , consider the polynomial
where denotes the curvature tensor of . Observe that we can formally extend the polynomial to every (even when does not make sense for ). Moreover, from the explicit formula for , which can be found in the Appendix A.2, we see that fixing , the map defines a linear function on . Thus, it is enough to prove that the polynomials and are non-negative. We will use LemmaΒ 4.5 to prove the stronger statement that and are polynomial sums of squares. Since is obtained from computing sectional curvatures, every monomial with non-trivial coefficient in satisfies . Hence, we do not lose generality replacing in Lemma 4.5 with
This change substantially reduces the size of the system (4.3) from to . Now we are looking for symmetric positive semi-definite matrices such that
This is a convex optimization problem, which, thanks to the reduction of the dimension mentioned above, can be successfully solved by the Python solver CVXOPT. We implemented the computer code in SageMath through two instances of SemidefiniteProgram(). However, this only yields numerical solutions, and since the condition of being positive semi-definite is a closed one, an exact solution is not guaranteed. Nonetheless, since the polynomial sums of squares are dense in the set of non-negative polynomials, exact solutions are expected to exist. Moreover, since has relatively few non-trivial coefficients, one can expect to find sparse solutions and . This is indeed the case, since rounding the numerical solutions lead us to the exact solution described as follows.
Define the index subsets
and let be the symmetric matrix defined as
Now it is routine to verify that is positive semi-definite and satisfies . Similarly, if
then
defines a symmetric positive semi-definite matrix satisfying . This concludes the proof of the theorem. β
Remark 4.6.
We are not certain if is a polynomial sum of squares for . Although it is not needed in the proof of Theorem 4.4, it would be interesting to know if this is indeed the case.
Remark 4.7.
The definition (4.1) of the metric resembles the construction of Berger spheres from the Hopf fibration by shrinking the metrics along the fibers. Moreover, if we restrict to , then is a Berger sphere. Recall that has (strictly) positive sectional curvature if and only if .
Appendix A
A.1. Standard zero divisors
In this appendix, we include Table 1 with the standard zero divisors of the sedenion algebra.
A.2. Expression for the sectional curvature
We write down the polynomial used in the proof of Theorem 4.4.
References
- [Arv03] A.Β Arvanitogeorgos, An introduction to Lie groups and the geometry of homogeneous spaces, Student Mathematical Library, no. v. 22, American Mathematical Society, Providence, RI, 2003.
- [BDI08] D.Β K. Biss, D.Β Dugger, and D.Β C. Isaksen, Large annihilators in CayleyβDickson algebras, Comm. Algebra 36 (2008), no.Β 2, 632β664.
- [BH87] A.Β Back and W.Β Y. Hsiang, Equivariant geometry and Kervaire spheres, Trans. Amer. Math. Soc. 304 (1987), no.Β 1, 207β227.
- [BLS20] T.Β C. Burness, M.Β W. Liebeck, and A.Β Shalev, The length and depth of compact Lie groups, Math. Z. 294 (2020), no.Β 3-4, 1457β1476.
- [CLR95] M.Β D. Choi, T.Β Y. Lam, and B.Β Reznick, Sums of squares of real polynomials, K-Theory and Algebraic Geometry: Connections with Quadratic Forms and Division Algebras. Summer Research Institute on Quadratic Forms and Division Algebras, July 6-24, 1992, University of California, Santa Barbara, CA (USA), Providence, RI: American Mathematical Society, 1995, pp.Β 103β126.
- [DZ79] J.Β E. DβAtri and W.Β Ziller, Naturally reductive metrics and Einstein metrics on compact Lie groups, Memoirs of the American Mathematical Society, no. 215, American Mathematical Society, Providence, RI, 1979.
- [GG19] A.Β B. Gillard and N.Β G. Gresnigt, Three fermion generations with two unbroken gauge symmetries from the complex sedenions, Eur. Phys. J. C 79 (2019), no.Β 5, 11 p.
- [Jen73] G.Β R. Jensen, Einstein metrics on principal fibre bundles, J. Differential Geom. 8 (1973), no.Β 4, 599β614.
- [Ker98] M.Β M. Kerr, New examples of homogeneous Einstein metrics, Michigan Math. J. 45 (1998), no.Β 1, 115β134.
- [Mor98] G.Β Moreno, The zero divisors of the Cayley-Dickson algebras over the real numbers, Bol. Soc. Mat. Mex., III. Ser. 4 (1998), no.Β 1, 13β28.
- [Oni92] A.Β L. Onishchik, The group of isometries of a compact Riemannian homogeneous space, Differential Geometry and Its Applications. Proceedings of a Colloquium, Held in Eger, Hungary, August 20-25, 1989, Organized by the JΓ‘nos Bolyai Mathematical Society, Amsterdam: North-Holland Publishing Company; Budapest: JΓ‘nos Bolyai Mathematical Society, 1992, pp.Β 597β616.
- [OR13] C.Β Olmos and S.Β Reggiani, A note on the uniqueness of the canonical connection of a naturally reductive space, Monatsh. Math. 172 (2013), no.Β 3-4, 379β386.
- [ORT14] C.Β Olmos, S.Β Reggiani, and H.Β Tamaru, The index of symmetry of compact naturally reductive spaces, Math. Z. 277 (2014), no.Β 3-4, 611β628.
- [PW98] V.Β Powers and T.Β WΓΆrmann, An algorithm for sums of squares of real polynomials, J. Pure Appl. Algebra 127 (1998), no.Β 1, 99β104.
- [Reg10] S.Β Reggiani, On the affine group of a normal homogeneous manifold, Ann. Global Anal. Geom. 37 (2010), no.Β 4, 351β359.
- [Reg24] by same author, Auxiliary computations for the article The Geometry of Sedenion Zero Divisors, https://github.com/silvioreggiani/sedenion-zero-div, 2024.
- [SA20] L.Β S. Saoud and H.Β Al-Marzouqi, Metacognitive Sedenion-Valued Neural Network and its Learning Algorithm, IEEE Access 8 (2020), 144823β144838.
- [Sag70] A.Β A. Sagle, Some Homogeneous Einstein Manifolds, Nagoya Math. J. 39 (1970), 81β106.
- [Zil07] W.Β Ziller, Examples of Riemannian manifolds with non-negative sectional curvature, Metric and Comparison Geometry. Surveys in Differential Geometry. Vol. XI., Somerville, MA: International Press, 2007, pp.Β 63β102.