As a direct continuation of Zwart [J. Math. Phys. 64(10), 101701 (2023)], which is built on the work of Müger and Tuset [Indagationes Math. 35(1), 114 (2024)], we reduce the Mathieu conjecture, formulated by Mathieu [Algèbra Non Commutative, Groupes Quantiques et Invariants, edited by Alex, J. and Cauchon, G. (Société Mathématique de France, Reims, 1997), Vol. 2, pp. 263–279], for Sp(N) and G2 to a conjecture involving functions over Rn×(S1)m with n,mN0. The proofs rely on Euler-style parametrizations of these groups, a specific version of the KAK decomposition, which we discuss and prove.

In 1997, Olivier Mathieu conjectured the following statement.

Conjecture 1.1

(Ref. 9). Let G be a compact connected Lie group. If f, h are complex valued finite-type functions such that GfPdg = 0 for all PN, then GfPh dg = 0 for all large enough P.

Just one year after the publication of Mathieu’s paper, Duistermaat and van der Kallen5 proved Mathieu’s conjecture for all abelian connected compact groups. Although the conjecture is still an open problem for any non-abelian group, some attempts were made. Dings and Koelink4 approached the conjecture for SU(2) by expressing the finite-type functions in terms of explicit matrix coefficients. Influenced heavily by this, Müger and Tuset10 reduced the Mathieu conjecture for SU(2) to a conjecture about certain Laurent polynomials. In 2023, this author proved,14 inspired by papers such as Refs. 2 and 12, that a similar reduction can be done for SU(N) and SO(N) to a conjecture regarding Laurent polynomials which also allow for complex square roots.

This paper is a direct continuation of Ref. 14 where we apply the same arguments to the infinite family of groups Sp(N) with NN and the exceptional group G2. As in Ref. 14, this can be achieved by using a generalization of the Euler decomposition which we will call the Euler angles. This decomposition is a more intricate version of the KAK decomposition, a well-known decomposition for connected Lie groups. The decomposition will be proven and applied to the groups Sp(N) and G2.

After having found the Euler angles, the finite-type functions of Sp(N) and G2 are considered, and we will reduce the Mathieu conjecture to a similar conjecture to that of Müger and Tuset in Ref. 10 and Zwart14 with a different weight function. To be more precise, any finite type function of Sp(N) or G2 reduces to a function f:Cn×RkC which can be written as f(z,x)=mcm(x)zm where m=(m1,,mk) is a multi-index where mij=1l1jZ for each i, for some lN, and cm(x) is a polynomial in x1, …, xk and 1x12,,1xk2. Assuming these functions satisfy certain conjectures, the Mathieu conjecture is proven for Sp(N) and G2. In Sec. II we will focus on the group Sp(N), while in Sec. III the group G2 will be considered. The final part of the paper is dedicated to proving the generalized Euler angles we use throughout this paper, with the corresponding explicit description of the Haar measure in this parametrization.

In this paper we will discuss Mathieu’s conjecture on Sp(N) and G2 with NN. We start by recalling Mathieu’s conjecture. To do so, we first introduce the notion of a finite-type function.

Definition 2.1.
Let G be a compact Lie group. A function f:GC is called a finite-type function if it can be written as a finite sum of matrix components of irreducible continuous representations, i.e.,
where (πj, Vj) is an irreducible continuous representation of G, and ajEnd(Vj).

Conjecture 2.2

(The Mathieu Conjecture9). Let G be a compact connected Lie group. If f, h are finite-type functions such that GfPdg = 0 for all PN, then GfPh dg = 0 for all large enough P.

This section is dedicated to the Lie group Sp(N), while Sec. III is dedicated to G2. This document is a direct continuation of Ref. 14, and we assume the reader is familiar with the techniques used there, for we will apply them throughout the paper.

In a similar fashion as in Ref. 14, we wish to apply some form of KAK decomposition to our groups in such a way that we can reduce the Mathieu conjecture. As already hinted in Ref. 14, the generalized Euler angles are an example of a more general decomposition, which is stated below in Theorem 2.3. This theorem is key to the rest of the paper, and will be applied to both Sp(N) and G2. We will prove the theorem in  Appendix A to keep this section focussed on Sp(N).

Theorem 2.3
(The Euler Angles Theorem). Let G be a simply connected compact Lie group with finite center, and let g be its Lie algebra. Let θ:gg be an involutive automorphism. Let k,p be the +1 and −1 eigenspace of θ, respectively, in such a way that g=kp. Fix a maximal abelian subalgebra ap and fix a set of (positive) roots Δp with respect to a. Then the mapping
(2.1)
is surjective and a diffeomorphism up to a measure zero set if we replace A by int(A). Here KG is the connected Lie group with Lie algebra k, and M=ZK(a). Here A is the closure of the connected component of the following set
in such a way that 0A. Here a+ is the closed positive Weyl chamber in a. In addition, the Haar measure decomposes in this parameterization as follows
for any measurable function h:GC, where C > 0 is a constant (independent of h), k2K with corresponding Haar measure dk2, k1k an arbitrary representative of k1MK/M with corresponding unique K-invariant measure dgK/M on K/M, and dH is the measure on a. Here J:AC is given by
(2.2)
where Δp+ is the set of positive roots on a.

The rest of the section will be dedicated to applying Theorem 2.3 to the group Sp(N). To do that, we first recall the definition of the compact symplectic group Sp(N).

Definition 2.4.
Let NN. We define the compact symplectic group as
where J=01N1N0 and 1N is the N × N identity matrix.

It is clear that the Lie algebra of Sp(N) is given by
(2.3)
Here, we denoted A as the adjoint of A, and U* the complex conjugate of U. In addition, since sp(N)u(2N), the Killing form is just given by the bilinear form ⟨X, Y⟩ = 4 Tr(XY). We note that the map
has the property θInt(sp(N)) and can be seen as a Cartan involution. It is easily seen that with this Cartan involution, we get the following decomposition:
(2.4)
(2.5)
which gives
One can take the following set as maximal abelian subalgebra:
where 0N is the N × N-matrix consisting of only zeros. With that, we can calculate M=ZK(a). Since MK, and it must commute with all options of aj, we see that the only options for xM is to be of the form
for some ϕjR. In specific, we see that
Hence ϕj = 0 or π for all j, giving
Therefore, applying Theorem 2.3, we find that
For the rest of this paper we will work with the latter diffeomorphism, since on U(N) some results regarding the Euler angle decomposition are already known.14 Let us consider the parametrization of U(N) first.
We remind ourselves of the Euler angle decomposition of SU(N) and on U(N). To describe this decomposition, we require an orthogonal basis for u(N). Let j = 1, 2, …, N − 1 and k = 1, 2, …, 2j and define the matrices λ0,λjsu(N) in the following way (in most physics papers the matrices {iλj}j are called Gell-Mann matrices, see e.g. Refs. 2, 12, and 13).

This is an orthonormal basis for u(N) as can easily be checked. With that, we get the following lemma.

Lemma 2.5
[Generalized Euler Angles SU(N)14]. Let N ≥ 1. Define inductively the mapping FSU(N):([0,π]×[0,2π]N2)×([0,π]×[0,2π]N3)××([0,π]×[0,2π])×[0,π]×0,π2N(N1)2×[0,2π]××0,2πN1SU(N) by FSU(1) ≡ 1 and
(2.6)
where A(k)(x,y)eλ3xeλ(k1)2+1y, and ψj0,π2,ωj0,2πj for all j. Here we denoted the product as
This mapping is surjective. Moreover it is a diffeomorphism on the interior of the hypercube onto its image which is SU(N) up to a measure zero set.

Corollary 2.6
[Generalized Euler Angles U(N)]. Let NN. Then the mapping FU(N):([0,π]×[0,2π]N2)×([0,π]×[0,2π]N3)××([0,π]×[0,2π])×[0,π]×0,π2N(N1)2×[0,2π]××0,2πN1×[0,2π]U(N) defined by
is a surjective map onto U(N), and is a diffeomorphism on the interior of the hypercube onto its image which is U(N) up to a measure zero set.

Proof.

This can be seen by the fact that SU(N) × U(1) ≃ U(N) as manifolds by sending (A, z) ↦ A · diag(z, 1, …, 1) with inverse AAdiag(det(A)1,1,,1),det(A). Using that xeix with x ∈ [0, 2π] is a diffeomorphism on the interior of the hypercube, and diag(z,1,,1)=eixλ0 in this parametrization, gives the corollary.□

Now that we know how to parametrize U(N), we go to U(N)/Z2N. To do so, note that the N rightmost matrices in FU(N)(ϕ1, …, ξ) are diagonal matrices. Since the symmetric space U(N)/Z2N is found by considering the coset elements gZ2N which is a finite set, and these sets are found by multiplying diagonal matrices on the right, we can argue that the ω1, …, ωN−1, ξ parameters will have a shorter range which will fully describe U(N)/Z2N. In other words we can see U(N)/Z2N as a submanifold of U(N). To be more precise, we note that
and so on. We conclude that we can find a parametrization of all elements in Z2N this way, hence giving the following parametrization.

Corollary 2.7
[Parametrization U(N)/Z2N]. Let NN. Then the mapping F̃U(N):([0,π]×[0,2π]N2)×([0,π]×[0,2π]N3)××([0,π]×[0,2π])×[0,π]×0,π2N(N1)2×[0,π]××0,πN1×[0,π]U(N) defined by
is a surjective map onto U(N)/Z2N, and is a diffeomorphism on the interior of the hypercube onto its image which is U(N)/Z2N up to a measure zero set.

With this corollary, we get the following lemma which parametrizes Sp(N). We will call this the Euler angles parametrization or the Euler angles decomposition of Sp(N).

Lemma 2.8
[Euler Angles Parametrization of Sp(N)]. Let NN and define G = Sp(N), fix a maximal abelian subalgebra ap and consider KU(N) as before. Then
diffeomorphic up to a measure zero set, where A=exp(A)exp(a) and M=ZK(a). Here
In addition, let FU(N) and F̃U(N) be as in Corollaries 2.6 and 2.7 respectively. Then we define the mapping FSp(N):([0,π]×[0,2π]N2)×([0,π]×[0,2π]N3)××([0,π]×[0,2π])×[0,π]×0,π2N(N1)2×[0,π]××0,πN1×[0,π]×A×([0,π]×[0,2π]N2)×([0,π]×[0,2π]N3)××([0,π]×[0,2π])×[0,π]×0,π2N(N1)2×[0,2π]××0,2πN1×[0,2π]Sp(N) by

This mapping is surjective onto Sp(N) and is a diffeomorphism on the interior of the hypercube onto its image which is Sp(N) up to a measure zero set.

Finally, the Haar measure dg is in this parametrization given by:
where C > 0 is some constant, dgK(ϕ1, …, ξ) is the Haar measure on KU(N) integrating over the variables ϕ1, …, ξ, and
Here the integration over the yj variables goes over exactly the area A.

Remark 2.9.

Note that we chose the variables that go over the symmetric space K/M to have a ∼ above their letters for bookkeeping, while the variables going over elements in K or A are without a ∼. This notation is kept throughout the paper.

Proof.
The first equation is proven by applying Theorem 2.3 to our case of Sp(N). Secondly, by applying Corollaries 2.6 and 2.7 to the first equation, we find that FSp(N)(ϕ̃,,ξ) is a surjective map onto Sp(N) and is a diffeomorphism up to a measure zero set. The only thing to prove is the area A. To do so, we recall the reduced root system of sp(N) with respect to a. Let us define the basis of a as the set {ej}j=1,,N where
Define the linear functional αja* as
Then sp(N)C has a root decomposition, with roots given by
We choose the set of positive roots to be
Then it is immediate that
where we wrote
Thus A can be described as
Note that if 0yjπ2 then yj ± ykπ automatically, so we can drop most inequalities, and only 0 ≤ yjyk adds new information. In other words, we can describe A as
which yields the theorem. We finally note that the measure on K/M can be seen to be the Haar measure on K times some positive constant. This can be seen by noting that up to a measure zero set, we can embed K/M into K. This allows us to note that the left-invariant measure on K/M can then be seen as the measure on K as well up to a C-function. But since both dgK/M and dgK are left-invariant, this a C function should be constant.□

Now that we have a suitable parametrization for Sp(N), we go to the Mathieu conjecture. We work in the same way as in Ref. 14. By Procesi (Ref. 11, Theorem 8.2.3), all finite-type functions of compact groups GGL(N,C) are generated by the matrix entries of G and the inverse of the determinant. With our parametrization of Sp(N) in Theorem 2.8, we can thus describe all finite-type functions almost everywhere. Together with the fact that the determinant of all matrices in Sp(N) is one, we see that any finite type function on Sp(N) can be written as a sum of products of the matrix entries of Sp(N), which is by Theorem 2.8 is described by the matrix entries of K/M, expA and K. Since the matrix entries of K (and of K/M) are by Corollary 2.6 given by the matrix entries of SU(N) times the determinant of U(N), which again can be interpreted as finite-type functions on SU(N) times the determinant of U(N), we get that a finite type function f on Sp(N) are can be written as
(2.7)
for almost all gG, where we identified g=FSp(N)(ϕ̃1,,ξ) such that the parameters ϕ̃1,,ξ are well-defined, fijSU(N) and hijSU(N) are finite-type functions on SU(N), mij,nijZ and
where pijnN0 for all n = 1, 2, …, N and qijm{0,1} for all m = 1, 2, …, N. We sum over two separate indices to make sure we have all the possible terms and allow for different powers of each term.

For more details regarding the finite-type functions on SU(N), we refer to Ref. 14. In a similar way as in Ref. 14, we have the following proposition.

Proposition 2.10.
Let f be a finite-type function on Sp(N) as in Eq. (2.7). Then for any PN we have
where
and
Here C′ > 0 is some constant and hijSU(N)̃ and JSU(N) are as in Ref. 14, Lemma 2.7. The function JSp(N) is defined as in Theorem 2.8, the set S* ≔ S1\{1} and the function f̄ijSU(N) is defined as
and

Proof.
We note that
To complete the lemma, we consider each integral separately. Using Ref. 14, Lemma 2.7, we immediately can claim that the last integral equals to
where hijSU(N)̃ is defined as in Ref. 14, Lemma 2.7, where we recall S* = S1\{1}. The same arguments can be applied to the first integral, only note that the ω̃j go over the range [0,πj] instead of [0,2πj] which gives a factor of 1/2 per parameter. Therefore, we get that the first integral can be written as
where
Finally note that
Note that
So if we subsitute ξj = sin(yj), we see that
Putting everything together gives:
Defining
and defining C=C(1)N(N+1)22(N1)2 gives the desired result.□

We note that Proposition 2.10 gives a sum of products of (possible roots of) polynomials, in a similar way as in Ref. 14. Let us give a name for these kind of polynomial.

Definition 2.11.
Let k,l,NN and f:[0,1]k×(S*)lC. We say f is a 1N-admissible function if f can be written as
where m=(m1,,ml) is a multi-index where mij=1N1jZ, and cm(x)C[x1,(1x12)1/2,,xk,(1xk2)1/2] is a complex polynomial in xi and 1xi2. We define the collection of m for which cm0 the spectrum of f, and it will be denoted by Sp(f).

Remark 2.12.

Note that in Ref. 14, we called these functions SU(N)-admissible functions.

It is clear from Proposition 2.10 that f̂ is a 1N-admissible function. Motivated by Refs. 10 and 14, we make the following conjecture.

Conjecture 2.13.
Let f:[0,1]N2×(S*)N(N+1)C be a 1N-admissible function. If
for all PN, then 0 does not lie in the convex hull of Sp(f).

Theorem 2.14.

Assume Conjecture 2.13 is true. Then the Mathieu conjecture is true for Sp(N).

Proof.
Let f, h be finite-type functions such that GfP(g)dg = 0 for all PN. We are interested in GfP(g)h(g)dg. Since f, h are finite-type functions, by the previous discussion we can conclude f and h have the form of Eq. (2.7). By linearity of the integral we can assume h to be of the form
where cC, the functions h1, h3 are finite-type functions on SU(N), m,nZ and
where KjN0 and Lj ∈ {0, 1}. Then the integral is of the form
(2.8)
Assume to the contrary that Gf(g)Ph(g)dg ≠ 0 for infinitely many P. That means that there exists a set {βij}i,j such that the integrals in Eq. (2.8) are non-zero. In particular, it means that
(2.9)
(2.10)
(2.11)
Our goal now is to make extensive use of the Haar measure properties to conclude that 0Sp(f). Note that any element gSp(N) can be written as
as in Theorem 2.3, where xK is a representative of xM, eHA and kK, and the Haar measure can be written as
Since dgK is just the Haar measure on K, for any smooth function f:GC we have
for any hK. Choosing kK we see that kh, hkK, and therefore by previous argument we see that only Eq. (2.11) would change. Also note that dgK/M is just the Haar measure on K as well up to a factor. In the same way we thus see that
where we used the left- and right-invariance of dgK/M. Because of the diffeomorphism up to a measure zero in Theorem 2.3, we have that only Eq. (2.9) would change. Being able to change Eqs. (2.9) and (2.11) independently, we can conclude that Eqs. (2.9)(2.11) are all independent equations.
But note that the Eqs. (2.9) and (2.11) are statements of finite-type functions on U(N). Expanding the finite-type functions on SU(N), in the same line as in Ref. 14, gives
and
where fijSU(N1) and hSU(N−1) are defined inductively, and the indices satisfy
Applying the same steps as in the proof of Ref. 14, Theorem 2.11, we find that
for all m=1,,N(N1)2 and n = 1, …, N. For more details we refer to Ref. 14. Using the same arguments for f̃ijSU(N), assigning a tilde on the indices of f̃ijSU(N), we find that
for all m=1,,N(N1)2 and n = 1, …, N. Finally, we note that
which must be equivalent to the map ξ̃ξ̃+t. This mapping can only be invariant for all tR if
In the same way we get
which gives in the same way
In other words, adding it all together, we get that
Multiplying both sides with 1P and letting P go to zero, gives a sequence that converges to 0. Since the convex hull is a closed set, 0 ∈ Conv(Sp(f)). But this is a contradiction with Conjecture 2.13. So Sp(N)f(g)Ph(g)dg ≠ 0 for only finitely many P, so Sp(N)f(g)Ph(g)dg = 0 for P large enough.□

Next, we shift our focus to G2. The structure of this section will be the same as before: we first find an Euler angles decomposition and then translate the Mathieu conjecture to a conjecture on [0,1]m×(S1)n for some n,mN. However, G2 is not by definition defined as a closed subgroup of GL(n,C) for some n. Although that is not a necessity to apply Theorem 2.3, we will do so to make much of the calculations actually computable. For that, we make extensive use of Ref. 3 which gives an explicit embedding of G2 into GL(7,R) and a way to do the Euler angles decomposition.

The group G2 can be seen as the set of automorphisms on the octonions.1 Let O be the set of octonions. We can see O as a 8-dimensional vectorspace over R, the spanning vectors being e0, …, e7 where e0 corresponds to the real unit, and ej to imaginary units for all j = 1, …, 7. There exists a natural multiplication on O which we will not need for our purpose. For more information on the octonions we refer to Ref. 1. However, it is known1 that G2 can be seen as the automorphism group of O. Since A is invertible and is linear, it must leave e0 fixed and can only permute e1, …, e7. Therefore one can consider G2 as a closed subset of GL(7,R).

Since the Lie algebra of automorphisms on an algebra A is the set of derivations on A, the Lie algebra g2gl(7,R) should therefore be the set of derivations on O. One can compute these, and the generators are given in  Appendix B.

We note that G2 is a simply connected compact simple Lie group, hence we can apply Theorem 2.3. In the same way as in Sp(N), we first find a Cartan involution. We choose the analytic involutive automorphism given by
It is clear that θInt(g2) and one can easily see that only matrices of the form A00B, where Agl(3,R) and Bgl(4,R), are invariant under θ. The Lie algebra k is then given by
where k1spanR(λ1,λ2,λ3) and k2spanR(λ8,λ9,λ10), both isomorphic to the simple Lie algebra su(2) and [k1,k2]=0. Note that the matrices {λj}j are defined in  Appendix B. This defines our K to be
Since [k1,k2]=0 one would expect K to be isomorphic to SU(2) × SU(2). Let us define the latter set by
However, we note that KK̂ since the group homomorphism
has ker f = {(17, 17), (diag(13, −14), diag(13, −14))}. So we see that
where Z2kerf and K̂ is the double cover of K.

So to apply Theorem 2.3, let us first describe the Euler angle decomposition of K. The Euler angle decomposition on K̂ can be described as follows.

Lemma 3.1.
Let K̂ be as before, then up to a measure zero set there is a diffeomorphism such that
as manifolds. Here

Proof.
The fact that C can be seen as R2, where multiplication with i can seen as applying the matrix
we see that SU(2) can be readily been embedded into SO(4), and we know that SU(2) can be described by the Euler angle decomposition as in, for example Refs. 4, 10, and 14, by
where
and
Applying this to every component gives the result.□

This way, we see that if kK̂ then
where ϕ1, ϕ2 ∈ [0, π], ψ1, ψ2 ∈ [0, π/2] and ω1, ω2 ∈ [0, 2π]. Note that eπλ3=diag(13,14), and since ek1ek2=ek2ek1, we can pull the mod Z2 operation into the first component, i.e., we get a unique description of gK up to a measure zero set if we restrict the range of ω1 to ω1 ∈ [0, π]. So we see that the following lemma holds.

Lemma 3.2
(Euler angle decomposition of K). Let K be as before. Define the mapping FK:[0,π]×0,π2×[0,π]2×[0,π2]×[0,2π]K by
Then this mapping is a smooth surjective map, and is a diffeomorphism up to a measure zero set.

With the Euler angle decomposition of K found, we note that the pair (G2, K) is the Riemannian symmetric pair associated with (g2,θ). To get the rest of the data we need for Theorem 2.3, we need to find a maximal abelian subalgebra ap, where
We choose the maximal abelian subalgebra to be
To apply Theorem 2.3, we need to calculate M. By definition, xM=ZK(a) if and only if xK and
for all Ha. A direct computation shows that, if we define
then M = {1, σ, η, ση}. With that, we get the following lemma.

Lemma 3.3
(Euler Angle decomposition of G2). Consider the compact simple Lie group G2. Let K=ek be as before. Then
diffeomorphic up to a measure zero set, where A=expA and M = {1, σ, η, ση} as before. In addition, define the mapping
by
where we repeat ϕ̃1,ϕ̃2,ϕ1,ω1,ϕ2[0,π], ψ̃1[0,π4], ω̃1,ψ̃2,ψ1,ψ2[0,π2] and ω̃2,ω2[0,2π] and y1,y2A for clarity. Here
Then FG2 is surjective and, up to a measure zero set, a diffeomorphism onto G2.
Finally, the Haar measure dg decomposes as follows
where C > 0 is some constant, dgK is the Haar measure on K, dgK/M the unique left-invariant measure on K/M, both given by
(3.1)
(3.2)
and
Here the integration over the yj variables goes exactly over the area A.

Proof.
The first part of the Theorem is proven by just applying Theorem 2.3 to G2. To prove the claims for FG2, we need to find a parametrization on K, on A and on K/M. The parametrization on K is given by Lemma 3.2, and the parametrization of A is immediate since a is abelian. So we only need to find a parametrization of K/M. We recall that M is discrete and finite, so we can describe K/M as a subset of K. So let gK. Using Lemma 3.2 to get the decomposition g=eϕ1λ3eψ1λ2eω1λ3eϕ2λ8eψ2λ9eω2λ8, we see immediately that
In the same way, after some calculations, we see that
Note we need to decompose this more, for in our decomposition ψj[0,π2] and is not described by this range. However, note that
and the similarly for eψ2λ9. In other words, we see that
Therefore, the equivalence relation on K to go to K/M is given by
So if we reduce our interval of ω1 to ω10,π2 and the interval of ψ1 to ψ10,π4, we can describe kMK/M explicitly by
for some unique ϕ1, ϕ2, ∈ [0, π], ψ1[0,π4], ω1,ψ2[0,π2] and ω2 ∈ [0, 2π], up to a measure zero set.
Now we are in the position to apply Theorem 2.3, by applying both our Euler angle decompositions of K, as given in Lemma 3.2, and K/M which was given above. This gives that any gG2 can be written uniquely, up to a measure zero set, as
where ϕ̃1,ϕ̃2,ϕ1,ω1,ϕ2[0,π], ψ̃1[0,π4], ω̃1,ψ̃2,ψ1,ψ2[0,π2] and ω̃2,ω2[0,2π] and y1,y2A, where A is as in Theorem 2.3.
Note that both K and K/M can be seen as open subsets of SU(2) × SU(2), hence the Haar measure is given by the Haar measure on SU(2) × SU(2) restricted to K and K/M times some constant, i.e., there exist C′, C″ > 0 such that the Haar measure on K and the left-invariant measure on K/M are given by
proving the Haar measure decomposition. To get an explicit form of A, and the corresponding Haar measure on expA, we recall the root decomposition of g2 with respect to a. The details of the root decomposition are in  Appendix B. Recall a=span(λ5,λ11). So we define the linear functionals α,βa* as
As will be shown in  Appendix B, the functionals , are the simple roots, and the set of positive roots are given by
It is then immediate from Eq. (2.2) that for H = y1λ5 + y2λ11 we get
With that, we see that
Note that this means that 0yiπ2, hence y1 ± y2π does not give any new information. In the same way we see that 0 ≤ y1 − 3y2 gives that y213y1 which is also satisfied by 0 ≤ y1y2. Finally if y1π2 then y1+3y2π2+π2. In other words, we can equally describe A by
(3.3)
which concludes the proof.□

In this section, we will follow the same reasoning as in Sec. III B to reduce the Mathieu conjecture to a more abelian version. We note that G2SL(7,C), hence the finite-type functions of G2 are generated by the matrix entries and the inverse determinant. Since we have, up to a measure zero set, a description of the matrix entries using Lemma 3.3, we see that any finite type function f is given by, using the diffeomorphism g=FG2(ϕ̃1,,ω2),
(3.4)
Here kijpZ, lijpN0 and mijp{0,1} for all i, j, p. We reduced mijp to be in {0, 1} by using sin(x)2 + cos(x)2 = 1. In a similar way as in Sec. II B, we can rewrite the condition G2fPdg in the following way.

Proposition 3.4.
Let f be a finite-type function on G2 as in Eq. (3.4). Then for any PN we have
where C=C252>0 is some constant,
and
(3.5)
and finally

Proof.
The proof goes analoguous to the Proof of Proposition 2.10 or the proof of Ref. 14, Lemma 2.7. By Lemma 3.3 and Eq. (3.4) we have
We will follow the steps as in Ref. 14. In short, we use the multinomium theorem to expand the exponent of the sum of matrix components, and then we use the following equations
for p,q,k,lN0 with l > 0 to get to integrals over S* and [0, 1]. Then we use the multinomial of Newton again to get the desired result. Note that there are a few integrals of the form 0πeikxdx, one integral of the form 0π/2eikxdx and one of the form 0π/4cosk(x)sinl(x)dx. The first integral can be transformed to an integral over S* by subsituting u = 2x, which gives
In the same way, we see that a subsitution of v = 4x gives
Finally, the last integral can be transformed to
This allows us to transform all the integrals, which give
Finally we would like to change the integrals going over y1 and y2. Putting in the area we found for A, we get that
where
Note now that doing the subsitution ξ1 = sin(y1) is possible, with only the boundary of the ξ2 integral is not easily transformable. However, it follows that
Let us call the right-handed side S(sin(y1)), then we see that sin(y1/3) = S(sin(y1)) which is some function depending on sin(x). Therefore we can apply the subsitution, and we this get the final form
where
which proves the lemma after noting 4C227=C225=C.□

Remark 3.5.

One might wonder why we chose to write the dependencies of ϕ̃1,ω2 in f in the way we did, i.e., why we chose a sum over eikϕ̃1 instead of sinl(ϕ̃1)cosm(ϕ̃1) when the integrals that come up in Proposition 3.4 are completely different. For in the end, the choice of eikϕ̃1 over sinl(ϕ̃1)cosm(ϕ̃1) should not change the finite-type function. The idea is to write those parameters as eikx only if the Jacobian J is independent of x, and keep the rest as sinl(x)cosm(x). This way the subsitutions are easily done.

In the same spirit as in Sec. III B, we see that f̃G2 is a 14-admissible function. To solve the Mathieu conjecture on G2, we assume the following conjecture.

Conjecture 3.6.
Let f:[0,1]6×(S*)8C be a 14-admissible function. If
for all PN, then 0 does not lie in the convex hull of Sp(f).

Theorem 3.7.

Assume Conjecture 3.6 is true. Then the Mathieu conjecture is true for G2.

Proof.
Let f, h be finite-type functions such that GfP(g)dg = 0 for all PN. We are interested in GfP(g)h(g)dg. As discussed, f and h have the form of Eq. (3.4). By linearity we can again assume h to be of the form
Using the multinomial theorem, we get that G2fP(g)h(g)dg can be written as
(3.6)
Let us assume G2fP(g)h(g)dg0 for infinitely many P. Then there exists at least one set of non-negative integers {βij}i,j such that i,jβij = P and such that the integral in Eq. (3.6) is non-zero. Applying Proposition 3.4 to this integral, we thus get that
To conclude the proof, we will show that i,jβijkija+Ka=0 for all a = 1, …, 8 and then make use of Conjecture 3.6. To do this, we will make extensive use of the Haar measure properties, i.e., make use of
for any measurable function f and any kG2. We will denote this as that the functions ggk and the function gkg are invariant mappings. We make extensive use of our parametrization given in Lemma 3.3 to see what the invariances of these functions mean for the possible parameters of Eq. (3.6).
We note that sending g ↦ exp(3)g is an invariant mapping for all tR. Going over to our parametrization this is equivalent to FG2(ϕ̃1,ϕ̃2,,ω2)FG2(ϕ̃1+t,ϕ̃2,,ω2) by bijectivity of FG2 up to a measure zero set, which would be equivalent to stating ϕ̃1ϕ̃1+t is an invariant mapping for the integral given in Eq. (3.6). Since t was chosen arbitrarily, this can only be true if
The same argument holds that gg exp(8) is an invariant mapping. By the same arguments this is equivalent to ω2ω2 + t is an invariant mapping which again can only be true if
Next, we remember that
where dy1dy2 is the measure on A, dgK/M is the unique left-invariant measure on K/M and dk is the Haar measure on K. This means that if we parametrize gG2 as g = xak where xK/M, a=exp(H)exp(A) and kK, then for any measurable function f we have
for any hK. In other words, the mapping g = xakxahk for any hK is also an invariant mapping. This way, the mapping g = xakxa exp(3)k is equivalent to sending ϕ1ϕ1 + t which is invariant. This can only be the case in Eq. (3.6) if
Next, remember that KSU(2)×SU(2)/Z2. In other words, it is the product of two commuting groups. So
if k=eϕ1λ3eψ1λ2eω1λ3eϕ2λ8eψ2λ9eω2λ8. Thus by previous arguments we see that the mapping g = xakxa exp(8)k is equivalent to the map ϕ2ϕ2 + t being invariant, which can only be the case if
Finally, we note that the measure dgK/M = dk as well, hence the same arguments can be reused to argue that the mapping g = xakxhak for any hK is also an invariant mapping. Applying that to g = xakx exp(3)ak is then equivalent to sending ω̃1ω̃1+t which can only be the case if
In the same way, the mapping g = xakx exp(8)ak is equivalent to sending ω̃2ω̃2+t which should be invariant, meaning that
and the map gg exp(3) is invariant which is equivalent to sending ω̃1ω̃1+t which can only be true if
Finally, using the invariance of the mapping g ↦ exp(8)g being equivalent to ϕ̃2ϕ̃2+t being invariant, we can conclude that
In other words, we have that i,jβijkija+Ka=0 for all a = 1, …, 8. This means that
Diving both terms by PN gives
Taking the limit P gives that the left hand side goes to 0. Hence there exists a limiting sequence in Sp(f) that converges to 0. Since the convex hull is a closed set, it means 0 ∈ Conv(Sp(f)) which is a contradiction with out assumption. So it cannot be that G2fPh0 for infinitely many P, hence proving the theorem.□

The author would like to thank Michael Müger for the continued help with the project, and the valuable discussions. He also wishes to thank Erik Koelink for the suggestion of looking at these two specific groups, and the valuable support.

The author has no conflicts to disclose.

Kevin Zwart: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (lead); Project administration (lead); Resources (lead); Software (lead); Supervision (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

In this section, we prove Theorem 2.3. For completeness we restate the theorem. The proof is based on Refs. 7 and 8.

Theorem 2.3
(The Euler Angles Theorem). Let G be a simply connected compact Lie group with finite center, and let g be its Lie algebra. Let θ:gg be an involutive automorphism. Let k,p be the +1 and −1 eigenspace of θ, respectively, in such a way that g=kp. Fix a maximal abelian subalgebra ap and fix a set of (positive) roots Δp with respect to a. Then the mapping
(A1)
is surjective and a diffeomorphism up to a measure zero set if we replace A by int(A). Here KG is the connected Lie group with Lie algebra k, and M=ZK(a). Here A is the closure of the connected component of the following set
in such a way that 0A. Here a+ is the closed positive Weyl chamber in a. In addition, the Haar measure decomposes in this parameterization as follows
for any measurable function h:GC, where C > 0 is a constant (independent of h), k2K with corresponding Haar measure dk2, k1k an arbitrary representative of k1MK/M with corresponding unique K-invariant measure dgK/M on K/M, and dH is the measure on a. Here J:AC is given by
(A2)
where Δp+ is the set of positive roots on a.

Proof.
We define the mapping
We claim that this mapping is a surjection, and up to a measure zero set a diffeomorphism.
First we note that f is well-defined. Indeed, taking any two representatives k, lkM gives k = lm with mM. Then it is clear Ad(k)H = Ad(l)Ad(m)H = Ad(l)H. In addition, we note that f(kM, exp(H), l) = k exp(H)k−1l. According to Helgason (Ref. 7, Theorem VII.8.6)
and if kexp(H)lKexp(A)K then H is unique. This show surjectivity of f. Now assume that
Then H1 = H2H as mentioned before, and so we are left with
We remember that θ:gg is an automorphism, and G is simply-connected, hence we can lift θ to an automorphism Θ: GG by Θ(exp(X)) = exp(θX). Then Θ(a) = a−1 for all aexpp and Θ(k) = k for all kK. Therefore we see that
The last equality sign shows that
where we defined kk21k1.

Let us now assume that α(H)πiZ for all αΔp, i.e., Hint(A). Using a similar argument as in Knapp (Ref. 8, Theorem 7.36) we get that the equality exp(Ad(k)2H) = exp(2H) requires Ad(k)H = H.

We claim that Zg(H)=ma where m=Zk(a). For if XZg(H), we see that XgC can be written as
where Xαgα and X0g0=mCaC. Then we note that
but ad(H)X = 0 by definition of X. So α(H)Xα = 0. However Ha+ so Xα = 0. This shows that Xma and so Zg(H)=ma. In particular Zp(H)=a. In this particular case it means that
So Ad(k)a=a and so kNK(a).

Since Ha+, using (Ref. 6, Lemma VII.2.2) there exists an sW(U, K) such that Ad(k)H = s · H, where W(U,K)NK(a)/ZK(a) is the analytic Weyl group. Since H lies in the positive Weyl chamber, the only reflection satisfying s · H = H is s = 1. So Ad(k) = 1, which means that kZK(a).

Going back to the definition of k, we thus see that k2 = k1k with kZK(a)=M. In other words k2M = k1M. It is them immediate that l1 = l2 and therefore we find that f is a bijection on int(A). Note that A and int(A) differ only by a measure zero set.

Note that f is smooth. This can be seen by first noting that the map ψ:K×Ag given by (k, H) ↦ Ad(k)H is smooth. Hence by the universal properties of the quotient, the map ψ̃:K/M×Ag, given by (kM, H) ↦ Ad(k)H, is also smooth. Therefore we see that the map expψ̃ is a smooth map, and multiplication is smooth, hence f is smooth.

So f:K/M×exp(int(A))×KG is bijective and smooth. We only need to show that the inverse is smooth, or that the tangent map is bijective everywhere, i.e., f is regular. We will show the latter. Let π: KK/M and ρ: GG/K the canonical maps. We remind ourselves that the map
is a regular map at every point (kM,H)K/M×int(A) (Ref. 7, Proposition VII.3.2). Now note that on int(A) the exponential exp:int(A)exp(int(A)) is a diffeomorphism. To prove this, note that since a is abelian, the differential of the exponent is just the identity. It is also injective, for if exp(H) = exp(H′) then exp(HH′) = e. This means Ad(exp(HH′)) is the identity mapping on g which in turn means α(H)α(H)2πiZ for all αΔp. If there exists an αΔp such that α(H) − α(H′) = 2πin with nZ{0} then note that either H or H′ is not in A since it means |α(H)| or |α(H′)| is bigger or equal than π which is impossible by definition of int(A). If α(H) − α(H′) = 0 for all αΔp then note that Δp is a root system, this means that Δp span a*, hence we must have HH′ = 0. So H = H′. Hence we conclude that
is a regular map at every point (kM,exp(H))K/M×exp(int(A)).
Note that Ad(k)Hp for all kK, and TeK(G/K)g/kp and thus TgK(G/K)=TeKτ(g)TeK(G/K) where τ(g): xKgxK is a diffeomorphism for all gG. This means that TgK(G/K)p for all gKG/K. Since Teρ:gg/kp by the same inclusion, we can conclude from this that the mapping
is a regular map. Now we note that
can be written as
where μ: G × GG, (g, h) ↦ gh is the multiplication map. In this way, we see that
which is clearly a bijection, since g=kp. In other words, we have that f is a smooth regular bijection, i.e., a diffeomorphism.
To get the decomposition of the Haar measure, we apply the techniques of Ref. 6, Sec. X.1.5 which gives that
where C > 0 is some constant. Here dk2 is the Haar measure on K, dH is the Haar measure on a and dkM is the unique left-invariant measure on K/M. We can rewrite this as
Using the fact that dk2 is a Haar measure, we can write this as
which proves the theorem.□

This section is based on Ref. 3. We can see G2 as the group of automorphisms on the octonions. This already shows that G2 can be embedded into GL(7,C), but we want to be explicit enough to make sure that the construction done in the main document can be verified explicitly. The basis elements for g2TeG2 are given by
Next, we briefly discuss the root decomposition of g2. It is clear that if we define the involution
that g2=kp where
and
In specific, note that [λ5, λ11] = 0. Choosing
gives a maximal abelian subalgebra in p. One can compute all commutators, and find that in the basis {λ1, …, λ4, λ6, …, λ10, λ12, …, λ14} the operators ad(λ5) and ad(λ11) look like
Computing the eigenvalues and eigenvectors of these matrices give the roots of g2. Let us define the following linear functionals α,βa* by
We claim and are the simple roots for the complex Lie algebra (g2)C. In fact, the root system is given by
as was expected. Choosing the positive roots as all the roots having +, we find that α, β are the simple roots. The root spaces are given by
1.
Baez
,
J. C.
, “
The octonions
,”
Bull. Am. Math. Soc.
39
(
2
),
145
205
(
2002
).
2.
Bertini
,
S.
,
Cacciatori
,
S. L.
, and
Cerchiai
,
B. L.
, “
On the Euler angles for SU(N)
,”
J. Math. Phys.
47
(
4
),
043510
(
2006
).
3.
Cacciatori
,
S. L.
and
Cerchiai
,
B. L.
, “
Exceptional groups, symmetric spaces and applications
,” arXiv:0906.0121 (
2009
).
4.
Dings
,
T.
and
Koelink
,
E.
, “
On the Mathieu conjecture for SU(2)
,”
Indagationes Math.
26
(
1
),
219
224
(
2015
).
5.
Duistermaat
,
J.
and
Van Der Kallen
,
W.
, “
Constant terms in powers of a Laurent polynomial
,”
Indagationes Math.
9
(
2
),
221
231
(
1998
).
6.
Helgason
,
S.
,
Differential Geometry and Symmetric Spaces
(
Academic Press, Inc.
,
1962
).
7.
Helgason
,
S.
,
Differential Geometry, Lie Groups and Symmetric Spaces
(
Academic Press, Inc.
,
1978
).
8.
Knapp
,
A. W.
,
Lie Groups Beyond an Introduction
, 2nd ed. (
Birkhäuser
,
Boston
,
2002
), ISBN: 0-8176-4259-5.
9.
Mathieu
,
O.
, “
Some conjectures about invariant theory and their applications
,” in
Algèbre Non Commutative, Groupes Quantiques et Invariants
, edited by
Alex
,
J.
and
Cauchon
,
G.
(
Société Mathématique de France
,
Reims
,
1997
), Vol.
2
, pp.
263
279
.
10.
Müger
,
M.
and
Tuset
,
L.
, “
The Mathieu conjecture for SU(2) reduced to an abelian conjecture
,”
Indagationes Math.
35
(
1
),
114
118
(
2024
).
11.
Procesi
,
C.
,
Lie Groups: An Approach Through Invariants and Representations
(
Springer
,
New York
,
2007
), ISBN: 978-0-387-28929-8.
12.
Tilma
,
T.
,
Byrd
,
M.
, and
Sudarshan
,
E.
, “
A parametrization of bipartite systems based on SU(4) Euler angles
,”
J. Phys. A: Math. Gen.
35
(
48
),
10445
10465
(
2002
).
13.
Tilma
,
T.
and
Sudarshan
,
E.
, “
Generalized Euler angle parametrization for SU(N)
,”
J. Phys. A: Math. Gen.
35
(
48
),
10467
10501
(
2002
).
14.
Zwart
,
K.
, “
On the Mathieu conjecture for SU(N) and SO(N)
,”
J. Math. Phys.
64
(
10
),
101701
(
2023
).