Skip to the Main Content

Note:These pages make extensive use of the latest XHTML and CSS Standards. They ought to look great in any standards-compliant modern browser. Unfortunately, they will probably look horrible in older browsers, like Netscape 4.x and IE 4.x. Moreover, many posts use MathML, which is, currently only supported in Mozilla. My best suggestion (and you will thank me when surfing an ever-increasing number of sites on the web which have been crafted to use the new standards) is to upgrade to the latest version of your browser. If that's not possible, consider moving to the Standards-compliant and open-source Mozilla browser.

December 7, 2025

Octonions and the Standard Model (Part 13)

Posted by John Baez

When Lee and Yang suggested that the laws of physics might not be invariant under spatial reflection — that there’s a fundamental difference between left and right — Pauli was skeptical. In a letter to Victor Weisskopf in January 1957, he wrote:

“Ich glaube aber nicht, daß der Herrgott ein schwacher Linkshänder ist.”

(I do not believe that the Lord is a weak left-hander.)

But just two days after Pauli wrote this letter, Chien-Shiung Wu’s experiment confirmed that Lee and Yang were correct. There’s an inherent asymmetry in nature.

We can trace this back to how the ‘left-handed’ fermions and antifermions live in a different representation of the Standard Model gauge group than the right-handed ones. And when we try to build grand unified theories that take this into account, we run into the fact that while we can fit the Standard Model gauge group into Spin(10)\text{Spin}(10) in various ways, not all these ways produce the required asymmetry. There’s a way where it fits into Spin(9)\text{Spin}(9), which does not work… and alas, this one has a nice octonionic description!

To keep things simple I’ll explain this by focusing, not on the whole Standard Model gauge group, but its subgroup SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3). Here is a theorem proved by Will Sawin in response to a question of mine on MathOverflow:

Theorem 10. There are exactly two conjugacy classes of subgroups of Spin(10)\text{Spin}(10) that are isomorphic to SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3). One of them has a representative that is a subgroup of Spin(9)Spin(10)\text{Spin}(9) \subset \text{Spin}(10), while the other does not.

I’ll describe representatives of these two subgroups; then I’ll say a bit about how they show up in physics, and then I’ll show you Sawin’s proof.

We can get both subgroups in a unified way! There’s always an inclusion

SO(m)×SO(n)SO(m+n) \text{SO}(m) \times \text{SO}(n) \to \text{SO}(m+n)

and taking double covers of each group we get a 2-1 homomorphism

Spin(m)×Spin(n)Spin(m+n) \text{Spin}(m) \times \text{Spin}(n) \to \text{Spin}(m+n)

In particular we have

Spin(4)×Spin(6)Spin(10) \text{Spin}(4) \times \text{Spin}(6) \to \text{Spin}(10)

so composing with the exceptional isomorphisms:

Spin(4)SU(2)×SU(2),Spin(6)SU(4) \text{Spin}(4) \cong \text{SU}(2) \times \text{SU}(2), \qquad \text{Spin}(6) \cong \text{SU}(4)

we get a 2-1 homomorphism

k:SU(2)×SU(2)×SU(4)Spin(10) k \colon \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \to \text{Spin}(10)

Now, there are three obvious ways to include SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) in SU(2)×SU(2)×SU(4)\text{SU}(2) \times \text{SU}(2) \times \text{SU}(4). There is an obvious inclusion

j:SU(3)SU(4) j \colon \text{SU}(3) \hookrightarrow \text{SU}(4)

but there are three obvious inclusions

,r,δ:SU(2)SU(2)×SU(2) \ell, r, \delta \colon \text{SU}(2) \hookrightarrow \text{SU}(2) \times \text{SU}(2)

namely the left one:

:SU(2) SU(2)×SU(2) g (g,1) \begin{array}{ccc} \ell \colon \text{SU}(2) &\to& \text{SU}(2) \times \text{SU}(2) \\ g & \mapsto & (g,1) \end{array}

the right one:

r:SU(2) SU(2)×SU(2) g (1,g) \begin{array}{ccc} r \colon \text{SU}(2) &\to& \text{SU}(2) \times \text{SU}(2) \\ g & \mapsto & (1,g) \end{array}

and the diagonal one:

δ:SU(2) SU(2)×SU(2) g (g,g) \begin{array}{ccc} \delta \colon \text{SU}(2) &\to& \text{SU}(2) \times \text{SU}(2) \\ g & \mapsto & (g,g) \end{array}

Combining these with our earlier maps, we actually get a one-to-one map from SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) to Spin(10)\text{Spin}(10). So we get three subgroups of Spin(10)\text{Spin}(10), all isomorphic to SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3):

  • There’s the left subgroup G G_\ell, which is the image of this composite homomorphism:

SU(2)×SU(3)×jSU(2)×SU(2)×SU(4)Spin(4)×Spin(6)kSpin(10) \text{SU}(2) \times \text{SU}(3) \stackrel{\ell \times j}{\longrightarrow} \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \cong \text{Spin}(4) \times \text{Spin}(6) \stackrel{k}{\longrightarrow} \text{Spin}(10)

  • There’s the diagonal subgroup G δG_\delta, which is the image of this:

SU(2)×SU(3)δ×jSU(2)×SU(2)×SU(4)Spin(4)×Spin(6)kSpin(10) \text{SU}(2) \times \text{SU}(3) \stackrel{\delta \times j}{\longrightarrow} \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \cong \text{Spin}(4) \times \text{Spin}(6) \stackrel{k}{\longrightarrow} \text{Spin}(10)

  • And there’s the right subgroup G rG_r, which is the image of this:

SU(2)×SU(3)r×jSU(2)×SU(2)×SU(4)Spin(4)×Spin(6)kSpin(10) \text{SU}(2) \times \text{SU}(3) \stackrel{r \times j}{\longrightarrow} \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \cong \text{Spin}(4) \times \text{Spin}(6) \stackrel{k}{\longrightarrow} \text{Spin}(10)

The left and right subgroups are actually conjugate, but the diagonal one is truly different! We’ll prove this by taking a certain representation of Spin(10)\text{Spin}(10), called the Weyl spinor representation, and restricting it to those two subgroups. We’ll get inequivalent representations of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3). This proves the two subgroups aren’t conjugate.

This argument is also interesting for physics. When restrict to the left subgroup, we get a representation of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) that matches what we actually see for one generation of fermions! This is the basis of the so-called SO(10)\text{SO}(10) grand unified theory, which should really be called the Spin(10)\text{Spin}(10) grand unified theory.

(In fact this works not only for SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) but for the whole Standard Model gauge group, which is larger. I’m focusing on SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) just because it makes the story simpler.)

When we restrict the Weyl spinor representation to the diagonal subgroup, we get a representation of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) that is not physically correct. Unfortunately, it’s the diagonal subgroup that shows up in several papers connecting the Standard Model gauge group to the octonions. I plan to say a lot more about this later.

The left subgroup

Let’s look at the left subgroup G G_\ell, the image of this composite:

SU(2)×SU(3)×jSU(2)×SU(2)×SU(4)Spin(4)×Spin(6)kSpin(10) \text{SU}(2) \times \text{SU}(3) \stackrel{\ell \times j}{\longrightarrow} \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \cong \text{Spin}(4) \times \text{Spin}(6) \stackrel{k}{\longrightarrow} \text{Spin}(10)

Spin(10)\text{Spin}(10) has a 32-dimensional unitary representation called the ‘Dirac spinor’ representation. This representation is really on the exterior algebra Λ 5\Lambda \mathbb{C}^5. It’s the direct sum of two irreducible parts, the even grades and the odd grades:

Λ 5Λ even 5Λ odd 5 \Lambda \mathbb{C}^5 \cong \Lambda^{\text{even}} \mathbb{C}^5 \oplus \Lambda^{\text{odd}} \mathbb{C}^5

Physicists call these two irreducible representations ‘right- and left-handed Weyl spinors’, and denote them as 16\mathbf{16} and 16*\mathbf{16}\ast since they’re 16-dimensional and one is the dual of the other.

Let’s restrict the 16\mathbf{16} to the left subgroup G G_\ell and see what we get.

To do this, first we can restrict the 16\mathbf{16} along kk and get

214124* \mathbf{2} \otimes \mathbf{1} \otimes \mathbf{4} \; \oplus \; \mathbf{1} \otimes \mathbf{2} \otimes \mathbf{4}\ast

Here 1\mathbf{1} is the trivial representation of SU(2)\text{SU}(2), 2\mathbf{2} is the tautologous representation of SU(2)\text{SU}(2), and 4\mathbf{4} is the tautologous rep of SU(4)\text{SU}(4).

Then let’s finish the job by restricting this representation along ×j\ell \times j. Restricting the 4\mathbf{4} of SU(4)\text{SU}(4) to SU(3)\text{SU}(3) gives 31\mathbf{3} \oplus \mathbf{1}: the sum of the tautologous representation of SU(3)\text{SU}(3) and the trivial representation. Restricting 21\mathbf{2} \otimes \mathbf{1} to the left copy of SU(2)\text{SU}(2) gives the tautologous representation 2\mathbf{2}, while restricting 12\mathbf{1} \otimes \mathbf{2} to this left copy gives 11\mathbf{1} \oplus \mathbf{1}: the sum of two copies of the trivial representation. All in all, we get this representation of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3):

2(31)(11)(3*1) \mathbf{2} \otimes (\mathbf{3} \oplus \mathbf{1}) \; \oplus \; (\mathbf{1} \oplus \mathbf{1}) \otimes (\mathbf{3}\ast \oplus \mathbf{1})

This is what we actually see for one generation of left-handed fermions and antifermions in the Standard Model! The representation 31\mathbf{3} \oplus \mathbf{1} describes how the left-handed fermions in one generation transform under SU(3)\text{SU}(3): 3 colors of quark and one ‘white’ lepton. The representation 3*1\mathbf{3}\ast \oplus \mathbf{1} does the same for the left-handed antifermions. The left-handed fermions form an isospin doublet, giving us the 2\mathbf{2}, while the left-handed antifermions have no isospin, giving us the 11\mathbf{1} \oplus \mathbf{1}.

This strange lopsidedness is a fundamental feature of the Standard Model.

The right subgroup would work the same way, up to switching the words ‘left-handed’ and ‘right-handed’. And by Theorem 10, the left and right subgroups must be conjugate in Spin(10)\text{Spin}(10), because now we’ll see one that’s not conjugate to either of these.

The diagonal subgroup

Consider the diagonal subgroup G δG_\delta, the image of this composite:

SU(2)×SU(3)δ×jSU(2)×SU(2)×SU(4)Spin(4)×Spin(6)kSpin(10) \text{SU}(2) \times \text{SU}(3) \stackrel{\delta \times j}{\longrightarrow} \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \cong \text{Spin}(4) \times \text{Spin}(6) \stackrel{k}{\longrightarrow} \text{Spin}(10)

Let’s restrict the 16\mathbf{16} to G δG_\delta.

To do this, first let’s restrict the 16\mathbf{16} along k:SU(2)×SU(2)×SU(4)Spin(10)k \colon \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) \to \text{Spin}(10) and get

214124* \mathbf{2} \otimes \mathbf{1} \otimes \mathbf{4} \; \oplus \; \mathbf{1} \otimes \mathbf{2} \otimes \mathbf{4}\ast

as before. Then let’s restrict this representation along δ×j\delta \times j. The SU(3)\SU(3) part works as before, but what happens when we restrict 21\mathbf{2} \otimes \mathbf{1} or 12\mathbf{1} \otimes \mathbf{2} along the diagonal map δ:SU(2)SU(2)×SU(2)\delta \colon \text{SU}(2) \to \text{SU}(2) \times \text{SU}(2)? We get 2\mathbf{2}. So, this is the representation of G δG_\delta that we get:

2(31)2(3*1) \mathbf{2} \otimes (\mathbf{3} \oplus \mathbf{1}) \; \oplus \; \mathbf{2} \otimes (\mathbf{3}\ast \oplus \mathbf{1})

This is not good for the Standard Model. It describes a more symmetrical universe than ours, where both left-handed fermions and antifermions transform as doublets under SU(2)\text{SU}(2).

The fact that we got a different answer this time proves that G G_\ell and G δG_\delta are not conjugate in Spin(10)\text{Spin}(10). So to complete the proof of Theorem 10, we only need to prove

  1. Every subgroup of Spin(10)\text{Spin}(10) isomorphic to SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) is conjugate to G G_\ell or G δG_\delta.

  2. G δG_\delta is conjugate to a subgroup of Spin(9)Spin(10)\text{Spin}(9) \subset \text{Spin}(10), but G G_\ell is not.

I’ll prove 2, and then I’ll turn you over to Will Sawin to do the rest.

Why the diagonal subgroup fits in Spin(9)\text{Spin}(9)

Every rotation of n\mathbb{R}^n extends to a rotation of n+1\mathbb{R}^{n+1} that leaves the last coordinate fixed, so we get an inclusion SO(n)SO(n+1)\text{SO}(n) \hookrightarrow \text{SO}(n+1), which lifts to an inclusion of the double covers, Spin(n)Spin(n+1)\text{Spin}(n) \hookrightarrow \text{Spin}(n+1). Since we have exceptional isomorphisms

Spin(3)SU(2),Spin(4)SU(2)×SU(2) \text{Spin}(3) \cong \text{SU}(2), \qquad \text{Spin}(4) \cong \text{SU}(2) \times \text{SU}(2)

it’s natural to ask how the inclusion Spin(3)Spin(4)\text{Spin}(3) \hookrightarrow \text{Spin}(4) looks in these terms. And the answer is: it’s the diagonal map! In other words, we have a commutative diagram

SU(2) Spin(3) δ SU(2)×SU(2) Spin(4) \begin{array}{ccc} \text{SU}(2) & \xrightarrow{\sim} & \text{Spin}(3) \\ \delta \downarrow & & \downarrow \\ \text{SU}(2) \times \text{SU}(2) & \xrightarrow{\sim} & \text{Spin}(4) \end{array}

Now, we can easily fit this into a larger commutative diagram involving some natural maps Spin(m)×Spin(n)Spin(m+n)\text{Spin}(m) \times \text{Spin}(n) \to \text{Spin}(m+n) and Spin(n)Spin(n+1)\text{Spin}(n) \to \text{Spin}(n+1):

SU(2) Spin(3) Spin(3)×Spin(6) Spin(9) δ SU(2)×SU(2) Spin(4) Spin(4)×Spin(6) Spin(10) \begin{array}{ccccccc} \text{SU}(2) & \xrightarrow{\sim} & \text{Spin}(3) & \to & \text{Spin}(3) \times \text{Spin}(6) & \to & \text{Spin}(9) \\ \delta \downarrow & & \downarrow & & \downarrow & & \downarrow \\ \text{SU}(2) \times \text{SU}(2) & \xrightarrow{\sim} & \text{Spin}(4) & \to & \text{Spin}(4) \times \text{Spin}(6) & \to & \text{Spin}(10) \end{array}

We can simplify this diagram using the isomorphism Spin(6)SU(4)\text{Spin}(6) \cong \text{SU}(4):

SU(2)×SU(4) Spin(9) δ×1 SU(2)×SU(2)×SU(4) Spin(10) \begin{array}{ccccccc} \text{SU}(2) \times \text{SU}(4) & \to & \text{Spin}(9) \\ \delta \times 1 \downarrow & & \downarrow \\ \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) & \to & \text{Spin}(10) \end{array}

and then we can use our friend the inclusion j:SU(3)SU(4)j \colon \text{SU}(3) \to \text{SU}(4):

SU(2)×SU(3) 1×j SU(2)×SU(4) Spin(9) δ×1 SU(2)×SU(2)×SU(4) Spin(10) \begin{array}{ccccccc} \text{SU}(2) \times \text{SU}(3) & \xrightarrow{1 \times j} & \text{SU}(2) \times \text{SU}(4) & \to & \text{Spin}(9) \\ & & \delta \times 1 \downarrow & & \downarrow \\ & & \text{SU}(2) \times \text{SU}(2) \times \text{SU}(4) & \to & \text{Spin}(10) \end{array}

This shows that the diagonal subgroup G δG_\delta of Spin(10)\text{Spin}(10) is actually a subgroup of Spin(9)\text{Spin}(9)!

Why the left subgroup does not fit in Spin(9)\text{Spin}(9)

The three-fold way is a coarse classification of irreducible representations of compact Lie group. Every such representation is of one and only one of these three kinds:

1) not self-dual: not isomorphic to its dual,

2a) Self-dual and orthogonal: isomorphic to its dual via an invariant nondegenerate symmetric bilinear form, also called an orthogonal structure,

2b) Self-dual and symplectic: isomorphic to its dual via an invariant nondegenerate antisymmetric bilinear form, also called a symplectic structure.

I’ve written about how these three cases are related to the division algebras ,\mathbb{C}, \mathbb{R} and \mathbb{H}, respectively:

But we don’t need most of this here. We just need to know one fact: when nn is odd, every irreducible representation of Spin(n)\text{Spin}(n), and thus every representation of this Lie group, is self-dual! In particular this is true of Spin(9)\text{Spin}(9).

Why does this matter? Assume the left subgroup G Spin(10)G_\ell \subset \text{Spin}(10) is a subgroup of Spin(9)\text{Spin}(9). When we restrict the Weyl spinor representation of Spin(10)\text{Spin}(10) to Spin(9)\text{Spin}(9) it will be self-dual, like every representation of Spin(9)\text{Spin}(9). Then when we restrict this representation further to SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) it must still be self-dual, since the restriction of a self-dual representation is clearly self-dual.

However, we know this representation is

2(31)(11)(3*1) \mathbf{2} \otimes (\mathbf{3} \oplus \mathbf{1}) \; \oplus \; (\mathbf{1} \oplus \mathbf{1}) \otimes (\mathbf{3}\ast \oplus \mathbf{1})

and this is not self-dual, since 1*1\mathbf{1}\ast \cong \mathbf{1} and 2*2\mathbf{2}\ast \cong \mathbf{2} but 3*3\mathbf{3}\ast \ncong \mathbf{3}.

So, it must be that G G_\ell is not a subgroup of Spin(9)\text{Spin}(9).

Proof of Theorem 10

To complete the proof of Theorem 10 we just need to see why there are just two conjugacy classes of subgroups of Spin(10)\text{Spin}(10) isomorphic to SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3). But in fact Will Sawin proved a stronger result! He was answering this question of mine:

Define the Standard Model gauge group to be S(U(2)×U(3))\text{S}(\text{U}(2) \times \text{U}(3)), the subgroup of SU(5)\text{SU}(5) consisting of block diagonal matrices with a 2×22 \times 2 block and then a 3×33 \times 3 block. (This is isomorphic to the quotient of U(1)×SU(2)×SU(3)\text{U}(1) \times \text{SU}(2) \times \text{SU}(3) by the subgroup of elements (α,α 3,α 2(\alpha, \alpha^{-3}, \alpha^2) where α\alpha is a 6th root of unity.)

Up to conjugacy, how many subgroups isomorphic to the Standard Model gauge group does Spin(10)\text{Spin}(10) have?

This question is relevant to grand unified theories of particle physics, as explained here:

This article focuses on one particular copy of S(U(2)×U(3))\text{S}(\text{U}(2) \times \text{U}(3)) in Spin(10)\text{Spin}(10), given as follows. By definition we have an inclusion S(U(2)×U(3))SU(5)\text{S}(\text{U}(2) \times \text{U}(3)) \hookrightarrow \text{SU}(5), and we also have an inclusion SU(5)Spin(10)\text{SU}(5) \hookrightarrow \text{Spin}(10) because for any nn we have an inclusion SU(n)SO(2n)\text{SU}(n) \hookrightarrow \text{SO}(2n), and SU(n)\text{SU}(n) is simply connected so this gives a homomorphism SU(n)Spin(2n)\text{SU}(n) \hookrightarrow \text{Spin}(2n).

However I think there is also an inclusion S(U(2)×U(3))Spin(9)\text{S}(\text{U}(2) \times \text{U}(3)) \hookrightarrow \text{Spin}(9), studied by Krasnov:

Composing this with Spin(9)Spin(10)\text{Spin}(9) \hookrightarrow \text{Spin}(10), this should give another inclusion S(U(2)×U(3))Spin(10)\text{S}(\text{U}(2) \times \text{U}(3)) \hookrightarrow \text{Spin}(10), and I believe this one is ‘truly different from’ — i.e., not conjugate to — the first one I mentioned.

So I believe my current answer to my question is “at least two”. But that’s not good enough.

Sawin’s answer relies on the 3-fold way. we embed SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) in Spin(10)\text{Spin}(10), we are automatically giving this group an orthogonal 10-dimensional representation, thanks to the map Spin(10)SO(10)\text{Spin}(10) \to \text{SO}(10). We can classify the possibilities.
He writes:

There are infinitely many embeddings. However, all but one of them is “essentially the same as” the one you studied as they become equal to the one you studied on restriction to SU(2)×SU(3)\text{SU}(2)\times \text{SU}(3). The remaining one is the one studied by Krasnov.

I follow the strategy suggested by Kenta Suzuki.

SU(3)\text{SU}(3) has irreducible representations of dimensions 1,3,3,6,8,6,10,101,3,3,6,8,6, 10, 10, and higher dimensions. The 1010-dimensional ones are dual to each other, as are the 66-dimensional ones, so they can’t appear. The 33-dimensional ones are dual to each other and can only appear together. So the only 1010-dimensional self-dual representations of SU(3)\text{SU}(3) decompose as irreducibles as 8+1+18+1+1, 3+3+1+1+1+13+3+1+1+1+1, or ten 11s. All of these are orthogonal because the 8-dimensional representation is orthogonal. However, the ten 11s cannot appear because then SU(3)\text{SU}(3) would act trivially.

A representation of SU(3)×SU(2)\text{SU}(3) \times \text{SU}(2) is a sum of tensor products of irreducible representations of SU(3)\text{SU}(3) and irreducible representations of SU(2)\text{SU}(2). Restricted to SU(3)\text{SU}(3), each tensor product splits into a sum of copies of the same irreducible representation. So SU(2)\text{SU}(2) can only act nontrivially when the same representation appears multiple times. Since the 3+33+3 is two different 33-dimensional representation, only the 11-dimensional representation can occur twice. Thus, our 10-dimensional orthogonal representation of SU(3)×SU(2)\text{SU}(3) \times \text{SU}(2) necessarily splits as either the 88-dimensional adjoint repsentation of SU(3)\text{SU}(3) plus a 22-dimensional orthogonal representation of SU(2)\text{SU}(2) or the 66-dimensional sum of standard and conjugate [i.e., dual] representations of SU(3)\text{SU}(3) plus a 44-dimensional orthogonal representation of SU(2)\text{SU}(2). However, SU(2)\text{SU}(2) has a unique nontrivial representation of dimension 22 and it isn’t orthgonal, so only the second case can appear. SU(2)\text{SU}(2) has representations of dimension 1,2,3,41,2,3,4 of which the 22 and 44-dimensional ones are symplectic and so must appear with even multiplicity in any orthogonal representation, so the only nontrivial 44-dimensional orthogonal ones are 2+22+2 or 3+13+1.

So there are two ten-dimensional orthogonal representations of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) that are nontrivial on both factors, those being the sum of two different 33-dimensional irreducible representations of SU(3)\text{SU}(3) with either two copies of the two-dimensional irreducible representation of SU(2)\text{SU}(2) or the three-dimensional and the one-dimensional irreducible representation of SU(2)\text{SU}(2). The orthogonal structure is unique up to isomorphisms, so these give two conjugacy classes of homomorphisms SU(2)×SU(3)SO(10)\text{SU}(2) \times \text{SU}(3) \to SO(10) and thus two conjugacy classes of homomorphisms SU(2)×SU(3)Spin(10)\text{SU}(2) \times \text{SU}(3) \to \text{Spin}(10). The first one corrresponds to the embedding you studied while only the second one restricts to Spin(9)\text{Spin}(9) so indeed these are different.

To understand how to extend these to S(U(2)×U(3))\text{S}(\text{U}(2) \times \text{U}(3)), I consider the centralizer of the representation within Spin(10)\text{Spin}(10). Since the group is connected, this is the same as the centralizer of its Lie algebra, which is therefore the inverse image of the centralizer in SO(10)\text{SO}(10). Now there is a distinction between the two examples because the example with irrep dimensions 3+3+2+23+3+2+2 has centralizer with identity component U(1)×SU(2)\text{U}(1) \times \text{SU}(2) while the example with irrep dimensions 3+3+3+13+3+3+1 has centralizer with identity component U(1)\text{U}(1). In the second case, the image of U(2)×U(3)\text{U}(2) \times \text{U}(3) must be the image of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) times the centralizer of the image of SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3), so this gives a unique example, which must be the one considered by Krasnov.

In the first case, we can restrict attention to a torus U(1)×U(1)\text{U}(1) \times \text{U}(1) in SU(2)×SU(2)\text{SU}(2) \times \text{SU}(2). The center of S(U(2)×U(3))\text{S}(\text{U}(2) \times \text{U}(3)) maps to a one-dimensional subgroup of this torus, which can be described by a pair of integers. Explicitly, given a two-by-two-unitary matrix AA and a three-by-three unitary matrix BB with det(A)det(B)=1\det(A) \det(B) =1, we can map to U(5)\text{U}(5) by sending (A,B)(A,B) to Aγ aBγ bA \gamma^a \oplus B \gamma^b where γ=det(A)=det(B) 1\gamma = \det (A) = \det(B)^{-1}, and then map from U(5)\text{U}(5) to SO(10)SO(10). This lifts to the spin group if and only if the determinant in U(5)\text{U}(5) is a perfect square. The determinant is γ 1+2a1+3b=γ 2a+3b\gamma^{ 1 + 2a - 1 + 3b} = \gamma^{2a+3b} so a lift exists if and only if bb is even.

The only possible kernel of this embedding is the scalars. The scalar A=λ 3I 2,B=λ 2I 3A = \lambda^3 I_2, B = \lambda^{-2} I_3 maps to λ 3+6aI 2λ 2+6bI 3\lambda^{3+ 6a} I_2 \oplus \lambda^{-2 + 6b} I_3 and so the kernel is trivial if and only if gcd(3+6a,2+6b)=1\gcd(3+6a,-2 + 6b)=1.

However, there are infinitely many integer solutions a,ba,b to gcd(3+6a,2a+6b)=1\gcd(3+6a,-2a+6b)=1 with bb even (in fact, a random aa and even bb works with probability 9/π 29/\pi^2), so this gives infinitely many examples.


  • Part 1. How to define octonion multiplication using complex scalars and vectors, much as quaternion multiplication can be defined using real scalars and vectors. This description requires singling out a specific unit imaginary octonion, and it shows that octonion multiplication is invariant under SU(3)\mathrm{SU}(3).
  • Part 2. A more polished way to think about octonion multiplication in terms of complex scalars and vectors, and a similar-looking way to describe it using the cross product in 7 dimensions.
  • Part 3. How a lepton and a quark fit together into an octonion — at least if we only consider them as representations of SU(3)\mathrm{SU}(3), the gauge group of the strong force. Proof that the symmetries of the octonions fixing an imaginary octonion form precisely the group SU(3)\mathrm{SU}(3).
  • Part 4. Introducing the exceptional Jordan algebra 𝔥 3(𝕆)\mathfrak{h}_3(\mathbb{O}): the 3×33 \times 3 self-adjoint octonionic matrices. A result of Dubois-Violette and Todorov: the symmetries of the exceptional Jordan algebra preserving their splitting into complex scalar and vector parts and preserving a copy of the 2×22 \times 2 adjoint octonionic matrices form precisely the Standard Model gauge group.
  • Part 5. How to think of 2×22 \times 2 self-adjoint octonionic matrices as vectors in 10d Minkowski spacetime, and pairs of octonions as left- or right-handed spinors.
  • Part 6. The linear transformations of the exceptional Jordan algebra that preserve the determinant form the exceptional Lie group E 6\mathrm{E}_6. How to compute this determinant in terms of 10-dimensional spacetime geometry: that is, scalars, vectors and left-handed spinors in 10d Minkowski spacetime.
  • Part 7. How to describe the Lie group E 6\mathrm{E}_6 using 10-dimensional spacetime geometry. This group is built from the double cover of the Lorentz group, left-handed and right-handed spinors, and scalars in 10d Minkowski spacetime.
  • Part 8. A geometrical way to see how E 6\mathrm{E}_6 is connected to 10d spacetime, based on the octonionic projective plane.
  • Part 9. Duality in projective plane geometry, and how it lets us break the Lie group E 6\mathrm{E}_6 into the Lorentz group, left-handed and right-handed spinors, and scalars in 10d Minkowski spacetime.
  • Part 10. Jordan algebras, their symmetry groups, their invariant structures — and how they connect quantum mechanics, special relativity and projective geometry.
  • Part 11. Particle physics on the spacetime given by the exceptional Jordan algebra: a summary of work with Greg Egan and John Huerta.
  • Part 12. The bioctonionic projective plane and its connections to algebra, geometry and physics.
  • Part 13. Two ways to embed SU(2)×SU(3)\text{SU}(2) \times \text{SU}(3) in Spin(10)\text{Spin}(10), and their consequences for particle physics.
Posted at December 7, 2025 4:04 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3625

1 Comment & 0 Trackbacks

Re: Octonions and the Standard Model (Part 13)

Hi, John,

Should this “division algebras ℂ,ℂ and ℍ, respectively” be ” division algebras ℝ, ℂ and ℍ, respectively”.

David

Posted by: David on December 19, 2025 10:00 AM | Permalink | Reply to this

Post a New Comment