The Mellin transform assigned to the convolution Poisson kernel on higher rank Riemannian symmetric spaces of the non-compact type is equal to the wave kernel. This makes it possible to determine the poles and to deduce the singular expansion of this kernel by using the zeta function techniques on compact and non-compact manifolds. As a consequence, we studied the harmonic sums associated with the wave kernel. In particular, we derived its asymptotic expansion near 0 according to the Mellin-converse correspondence rule.
Citation: Ali Hassani. Singular expansion of the wave kernel and harmonic sums on Riemannian symmetric spaces of the non-compact type[J]. AIMS Mathematics, 2025, 10(3): 4775-4791. doi: 10.3934/math.2025219
[1] | Ali Hassani . On the meromorphic continuation of the Mellin transform associated to the wave kernel on Riemannian symmetric spaces of the non-compact type. AIMS Mathematics, 2024, 9(6): 14731-14746. doi: 10.3934/math.2024716 |
[2] | Ling Zhu . Asymptotic expansion of a finite sum involving harmonic numbers. AIMS Mathematics, 2021, 6(3): 2756-2763. doi: 10.3934/math.2021168 |
[3] | F. Z. Geng . Piecewise reproducing kernel-based symmetric collocation approach for linear stationary singularly perturbed problems. AIMS Mathematics, 2020, 5(6): 6020-6029. doi: 10.3934/math.2020385 |
[4] | Harivan R. Nabi, Hajar F. Ismael, Nehad Ali Shah, Wajaree Weera . W-shaped soliton solutions to the modified Zakharov-Kuznetsov equation of ion-acoustic waves in (3+1)-dimensions arise in a magnetized plasma. AIMS Mathematics, 2023, 8(2): 4467-4486. doi: 10.3934/math.2023222 |
[5] | Thongchai Botmart, Soubhagya Kumar Sahoo, Bibhakar Kodamasingh, Muhammad Amer Latif, Fahd Jarad, Artion Kashuri . Certain midpoint-type Fejér and Hermite-Hadamard inclusions involving fractional integrals with an exponential function in kernel. AIMS Mathematics, 2023, 8(3): 5616-5638. doi: 10.3934/math.2023283 |
[6] | Tong Wu, Yong Wang . Super warped products with a semi-symmetric non-metric connection. AIMS Mathematics, 2022, 7(6): 10534-10553. doi: 10.3934/math.2022587 |
[7] | Dazhao Chen . Weighted boundedness for Toeplitz type operator related to singular integral transform with variable Calderón-Zygmund kernel. AIMS Mathematics, 2021, 6(1): 688-697. doi: 10.3934/math.2021041 |
[8] | Salim Bouzebda, Amel Nezzal, Issam Elhattab . Limit theorems for nonparametric conditional U-statistics smoothed by asymmetric kernels. AIMS Mathematics, 2024, 9(9): 26195-26282. doi: 10.3934/math.20241280 |
[9] | Jiaye Lin . Evaluations of some Euler-type series via powers of the arcsin function. AIMS Mathematics, 2025, 10(4): 8116-8130. doi: 10.3934/math.2025372 |
[10] | M. Hafiz Uddin, M. Ali Akbar, Md. Ashrafuzzaman Khan, Md. Abdul Haque . New exact solitary wave solutions to the space-time fractional differential equations with conformable derivative. AIMS Mathematics, 2019, 4(2): 199-214. doi: 10.3934/math.2019.2.199 |
The Mellin transform assigned to the convolution Poisson kernel on higher rank Riemannian symmetric spaces of the non-compact type is equal to the wave kernel. This makes it possible to determine the poles and to deduce the singular expansion of this kernel by using the zeta function techniques on compact and non-compact manifolds. As a consequence, we studied the harmonic sums associated with the wave kernel. In particular, we derived its asymptotic expansion near 0 according to the Mellin-converse correspondence rule.
In this paper, we pursue the study of the Mellin transform associated with the main evolution equations on general Riemannian d-dimensional non-compact symmetric spaces X=G/K of higher real rank initiated in [1] with the wave equation, where G denotes a semisimple Lie group, connected, non-compact, with a finite center, and K is a maximal compact subgroup of G. Here the version of the Mellin transform is defined as a function on the complex plane by the relation
f∗(σ)=1Γ(σ)∫∞0f(s)sσ−1ds, | (1.1) |
for some function f defined on the positive real axis, and Γ denotes the usual gamma function.
Generally, for some absolutely integrable function f on a finite interval (s1,s2),0<s1<s2<∞, which satisfies the following bounds
f(s)=O(s−c1)⏟s→0,f(s)=O(s−c2)⏟s→∞, | (1.2) |
where c1 and c2 are positive constants such that c1<c2, the Mellin transform f∗ exists and it is an analytic function on the vertical strip c1<Re(σ)<c2, called a fundamental strip. In some cases, this strip may extend to a half-plane (c1=−∞ or c2=+∞) or to the whole complex σ-plane (c1=−∞ and c2=+∞).
The Mellin transform is closely connected with the Fourier and Laplace transforms. A direct way to invert Mellin's transformation (1.1) is to start from Fourier's inversion theorem so that the original function f(s) under some conditions can be recovered from its Mellin transform by the inverse Mellin formula
f(s)=∫a+i∞a−i∞Γ(σ)f∗(σ)s−σdσ, | (1.3) |
where the integration is along a vertical line through c1<Re(σ)=a<c2.
As always, several questions arise around the convergence and analyticity of the integral (1.1) outside the fundamental strip c1<Re(σ)<c2. Classically, assume that f(s) admits as s→0 a finite asymptotic expansion of the form
f(s)=∑q∈I,Ifiniteαqsq+O(sc3),−c3<−q≤c1. | (1.4) |
Then for c1<Re(σ)<c2, the integral (1.1) can be decomposed as follows:
f∗(σ)=1Γ(σ)∫10f(s)sσ−1ds+1Γ(σ)∫∞1f(s)sσ−1ds=1Γ(σ)∫10(f(s)−∑q∈Iαqsq)sσ−1ds+1Γ(σ)∫10∑q∈Iαqsq+σ−1ds⏟=1Γ(σ)∑q∈Iαqσ+q+1Γ(σ)∫∞1f(s)sσ−1ds. |
The first integral on the right defines an analytic function on the strip (−c3,∞), and the second for all σ∈C, so that f∗(σ) is continuable to a meromorphic function on the strip (−c3,c2) with simple poles of residue αqΓ(q) at σ=−q, and no other singularities. The grouping of the poles appearing in the middle sum is called the singular expansion of f∗(σ) in the strip (−c3,c2) and is written
f∗(σ)≍1Γ(σ)∑q∈Iαqσ+q. | (1.5) |
In summary, there is a remarkable correspondence of coefficients in asymptotic expansion (1.4) of the original function f(s) and in the singular expansion of its Mellin transform f∗(σ) expressed by the rule
sq↦1Γ(σ)1σ+q. | (1.6) |
More generally, even if the asymptotic expansion of f(s) as s→0 contains terms of the form sq(lns)k, where the k are nonnegative integers, it is possible to obtain poles of order greater than 1 at σ=−q, since in the strip (−c3,c2)
(∂∂σ)k∫10sq+σ−1ds=(−1)kk!(σ+q)k+1. | (1.7) |
It should be noted that the reverse way is still valid. In other words, the singularities of the Mellin transform f∗(σ) can encode under certain conditions an asymptotic expansion of the original function f(s) as s→0. Namely, if f is a continuous function on (0,∞) with Mellin transform f∗(σ) having a meromorphic continuation to the extended strip (c4,c2) with a finite number of poles and a sufficiently fast decrease (say f∗(σ)=O(|σ|−r),r>1 as |σ|→∞) such that the singular expansion of f∗(σ) on (c4,c1) is written
f∗(σ)≍∑q,kcq,k(σ−q)k, | (1.8) |
then an asymptotic expansion of f(s) as s→0 takes the form
f(s)=∑q,kcq,k(−1)k−1(k−1)!s−q(lns)k+O(s−c4). | (1.9) |
So, as in the direct way, the following converse correspondence rule holds:
A(σ−q)k+1↦(−1)kk!s−q(lns)k. | (1.10) |
More elements of the theory of the Mellin transform can be found, for example, in [2,3,4].
By in [1, formula (1.13)], the author asserts that the zeta function ζG/K on the symmetric space X=G/K appears as a certain limit of the Mellin transform of the Poisson kernel on this space. However, in the general case, it is well known that
ζ(σ)=1Γ(σ)∫∞01es−1sσ−1ds,Re(σ)>1=1Γ(σ)∫∞0(∑k≥1e−ks)sσ−1ds=G∗ζ(σ), | (1.11) |
where Gζ(s)=∑k≥1g(ks),g(s)=e−s,s>0.
So, the singular behavior of the zeta function (which is well known) comes from the asymptotic development of the original function g(s)=e−s,s→0, or more precisely of the function Gζ(s)=∑k≥1g(ks), and the reverse way is still true as already indicated. Note that Gζ(s) presents a special case of sums of the form
G(s)=∑k≥1λkg(μks),s>0 | (1.12) |
called harmonic sums, which are highly useful in the theory of Dirichlet series [5]. The λk and μk are being interpreted as amplitudes and frequencies, respectively, and g(s) is called the base function. Those expressions arise in applications of combinatorial theory, especially in the evaluation of algorithms where the problem is to find the behavior of G(s) when s tends to 0 or infinity. Motivated by these facts, we propose to study the asymptotic of G(s), in the case where g(s)=e−bsp(s−it,x), where b>0 is a real parameter and (p(s−it,x))s>0,t>0,x∈X denotes the Poisson kernel on Riemannian symmetric space X=G/K and whose Mellin transform equals the wave kernel noted w(t,x,σ,b), σ∈C.
Let Δ be the Laplace-Beltrami operator on X=G/K. Using the inverse spherical Fourier transform, we consider the wave convolution kernel associated with Δ as a bi-K-invariant kernel on G expressed as follows:
w(t,x,σ,b)=1|W|∫a⋆eit√|λ|2+|ρ|2(b+√|λ|2+|ρ|2)−σφλ(x)|c(λ)|−2dλ, | (1.13) |
for t>0,x∈Gandσ∈C,such thatRe(σ)>d=dim(X), and a real parameter b>0, where a⋆ denotes a vector dual of a maximal abelian subspace a in the Cartan decomposition of the Lie algebra g of the group G. Here, φλ denotes the spherical function on G of index λ∈a⋆ and c(λ) is the Harish-Chandra c-function. |W| denotes the cardinality of the Weyl group W associated with the root system Σ, and ρ is the half sum of positive roots counted with their multiplicities relative to a.
The kernel w(t,x,σ,b) satisfies the wave equation ∂2∂t2w(t,x,σ,b)=Δw(t,x,σ,b), and its spherical Fourier transform is given by
ˆw(t,λ,σ,b)=eit√|λ|2+|ρ|2(b+√|λ|2+|ρ|2)−σ,λ∈a⋆, |
and it represents the convolution kernel of the operator W(t,σ,b)=eit√−Δ(b+√−Δ)−σ on the symmetric space X:
W(t,σ,b)f(x)=f⋆w(t,x,σ,b)=∫Gw(t,y−1x,σ,b)f(y)dy. |
According to the formula
a−σ=1Γ(σ)∫∞0e−uauσ−1du∀a>0, | (1.14) |
we easily write the integral expression (1.13) differently:
w(t,x,σ,b)=1Γ(σ)∫∞0e−bs(1|W|∫a⋆e−(s−it)√|λ|2+|ρ|2φλ(x)|c(λ)|−2dλ)⏟p(s−it,x)sσ−1ds=1Γ(σ)∫∞0e−bsp(s−it,x)⏟˜p(s−it,x)sσ−1ds=˜p∗(r−it,x)(σ), | (1.15) |
where p(s−it,x) denotes the bi-K-invariant convolution kernel of the Poisson operator Ps−it.
We show that the tools employed in the study of zeta functions on compact and non-compact manifolds [6,7,8,9,10,11,12,13] can be generalized to w(t,x,σ,b)=˜p∗(r−it,x)(σ). Using pointwise kernel estimates proved in [14,15,16], we shall see that w(t,x,σ,b) is a well-defined analytic function on Re(σ)>d, for every x∈X=G/K. We prove that w(t,x,σ,b) extends meromorphically to the entire complex plane C with a finite number of simple poles on the real line at the points (σj=d−j)j=0,⋯,d−1, by using the short-time asymptotic expansion of the Poisson kernel
p(s−it,x)∼cds−d∑j≥0αj(t,x)sj,s→0+, |
and residues given by
res(w(t,x,σ,b),σj=d−j)=cdΓ(d−j)˜αj(t,x,b),∀x∈X=G/K. |
By the fundamental correspondence (1.6), we deduce the singular expansion of w(t,x,σ,b)
w(t,x,σ,b)≍cdΓ(σ)d−1∑j=0˜αj(t,x,b)σ+(j−d),σ∈C,t>0,x∈X=G/K, |
where
˜αj(t,x,b)=j∑l=0(−1)j−lbj−l(j−l)!αl(t,x),j∈N0. |
As an application, we study the harmonic sums
G(t,x,s,b)=∑k≥1˜p(ks−it,x)=∑k≥1e−bksp(ks−it,x),s>0,x∈X=G/K. |
In particular, we prove that the Mellin transformation performs a separation principle between the base function ˜p(s−it,x) and the amplitude λk=1 and the frequencies μk=k. More precisely, we state
G∗(t,x,σ,b)=ζ(σ)w(t,x,σ,b),Re(σ)>d,t>0,x∈X=G/K, |
which allows us to establish the singular expansion of the function Γ(σ)G∗(t,x,σ,b)
Γ(σ)G∗(t,x,σ,b)≍cd˜αd−1(t,x,b)(σ−1)2+a−1σ−1+cdd−2∑j=0ζ(d−j)˜αj(t,x,b)σ−(d−j), |
where a−1=res(Γ(σ)G∗(t,x,σ,b),σ=1). This translates by the converse correspondence rule (1.10) into the asymptotic expansion of G(t,x,s,b) as s→0
G(t,x,s,b)∼1s(−cd˜αd−1(t,x,b)lns+a−1)+cdd−2∑j=0ζ(d−j)˜αj(t,x,b)sj−d. |
This paper is organized as follows: After recalling some basic notations and reviewing little aspects of Spherical-Fourier analysis on non-compact symmetric spaces in Section 2, we prove the extension of σ↦w(t,x,σ,b) to a meromorphic function on the entire complex plane, and we deduce its singular expansion in Section 3. We devote Section 4 to the study of harmonic sums assigned to the wave kernel. Finally, we give some conclusions and comparative remarks about similar and possible future results in Section 5.
We recall the basic notations of Fourier analysis on Riemannian symmetric spaces of the non-compact type. We refer to [17,18,19] for geometric properties and more details for harmonic analysis on these spaces.
Throughout this paper, the symbol A≲B between two positive expressions means that there is a positive constant C such that A≤CB.
Let G be a non-compact connected semisimple Lie group with a finite center and K a maximal compact subgroup of G. Let g (resp., k) be the Lie algebra of G (resp., K) and consider the Cartan decomposition g=k⊕p of g. Here, p is the (K-invariant) orthogonal complement of k in g with respect to the Killing form ⟨.,.⟩ of g. This form induces a K-invariant scalar product on p, and, hence, a G-invariant Riemannian metric on the symmetric homogeneous manifold X=G/K whose tangent space at the origin eK is naturally identified with p. Fix a maximal abelian subspace a in p and denote by a⋆ (respectively, a⋆C) the real (respectively, complex) vector dual of a. The Killing form of g induces a scalar product on a⋆ and a C-bilinear form on a⋆C. If λ,μ∈a⋆C, let Hλ∈aC be determined by λ(H)=⟨Hλ,H⟩,∀H∈aC, and put ⟨λ,μ⟩=⟨Hλ,Hμ⟩. We put |λ|=⟨λ,λ⟩1/2 for λ∈a⋆. Denote by Σ the root system of g relative to a. Let W be the Weyl group associated with Σ, and let mλ denote the multiplicity of the root λ∈Σ. In particular, if a+ denotes the positive Weyl chamber in a corresponding to some fixed set Σ+ of positive roots, we have the Cartan decomposition G=Kexp(¯a+)K of G, where ¯a+ denotes the closure of a+. Each element x∈G is written uniquely as x=k1A(x)k2. We denote by |x|=|A(x)| the norm defined on G. Let ρ=12∑λ∈Σ+mλλ and let d be the dimension of X, and D be the dimension at infinity of X. Consider g=k⊕a⊕n, the Iwasawa decomposition of g, and the corresponding Iwasawa decomposition G=Kexp(a)N of G, where N is the analytic subgroup of G associated to the nilpotent subalgebra n. Denote by H(x) the Iwasawa component of x∈G in a. There is a basic estimate for this component (see [17, p. 476], )
|H(x)|≲|x|, | (2.1) |
and it is called Iwasawa projection. Finally, let ℓ=dim(a), the real dimension of a. By definition, ℓ is the real rank of G.
We identify functions on X=G/K with functions on G, which are K-invariant on the right and, hence, bi-K-invariant functions on G with functions on X, K-invariants on the left. If f is a sufficiently regular bi-K-invariant function on G, then its spherical-Fourier transform is a function on a⋆C defined by
ˆf(λ)=∫Gf(x)φ−λ(x)dx,λ∈a⋆C, |
where φλ denotes the spherical function on G defined by
φλ(x)=∫Ke<iλ−ρ,H(xk)>dk,x∈G,λ∈a⋆C, |
and it satisfies the basic estimates
|φλ(x)|≤φ0(x),∀x∈G,∀λ∈a⋆C. |
The Plancherel and inversion formulas for the spherical transform are, respectively, given by
∫G|f(x)|2dx=1|W|∫a⋆|ˆf(λ)|2|c(λ)|−2dλ, |
f(x)=1|W|∫a⋆φλ(x)ˆf(λ)|c(λ)|−2dλ,x∈G, |
where |W| denotes the order of the Weyl group W and c(λ) is the Harish-Chandra c-function, and it satisfies the estimate
|c(λ)|−2≲(1+|λ|)d−ℓ, | (2.2) |
together with all its derivatives.
In this section, we use the following key pointwise estimates (see [14,15,16])
∀t>0,∀x∈X=G/K,|p(s−it,x)|≲{s−d,0<s≤1;s−D,s>1; | (3.1) |
to study the analyticity of the wave kernel w(t,x,σ,b) given by the rule:
w(t,x,σ,b)=1Γ(σ)∫∞0˜p(s−it,x)sσ−1ds=˜p∗(r−it,x)(σ), | (3.2) |
for a complex number σ∈C such that Re(σ)>d, and fixed x∈X=G/K and a real parameter b>0, where ˜p(s−it,x)=e−bsp(s−it,x) denotes the bi-K-invariant convolution kernel of the Poisson operator Ps−it.
Theorem 3.1. The function σ↦w(t,x,σ,b) extends meromorphically to C with simple poles located at σj=d−j;j∈N0, for all x∈X and for all positive real numbers t>0.
Proof. We divide the proof into three steps.
Step 1. The convergence of the integral given by Eq (3.2), where Re(σ)>d, is easily handled. Clearly, using the pointwise estimates in Eq (3.1), we have the following straightforward bound:
∫∞0|e−bsp(s−it,x)sσ−1|ds≲∫10s−d+Re(σ)−1ds+∫∞1e−bss−D+Re(σ)−1ds, |
and observe that the first integral converges for Re(σ)>d. However, the second integral is controlled as follows:
∫∞1e−bss−D+Re(σ)−1ds≲∫∞1e−bs/2ds<∞. |
This concludes the absolute convergence for Re(σ)>d.
Step 2. We check the holomorphy of ˜p∗(r−it,x)(σ)=w(t,x,σ,b). First, we split ˜p∗(r−it,x)(σ) into a local part ˜p∗0(r−it,x)(σ) and a part at infinity ˜p∗∞(r−it,x)(σ).
Write ˜p∗(r−it,x)(σ)=˜p∗0(r−it,x)(σ)+˜p∗∞(r−it,x)(σ), where
˜p∗0(r−it,x)(σ)=1Γ(σ)∫10˜p(s−it,x)sσ−1ds, | (3.3) |
and
˜p∗∞(r−it,x)(σ)=1Γ(σ)∫∞1˜p(s−it,x)sσ−1ds. | (3.4) |
To treat the local part, consider the vertical band
Sβ,γ={σ∈C;d<β<Re(σ)<γ},∀0<β<γ<∞, |
and let ε be a positive real number, and introduce the sequence (˜p∗0,ε(r−it,x)(σ))ε>0 of holomorphic functions on Sβ,γ
˜p∗0,ε(r−it,x)(σ)=1Γ(σ)∫1ε˜p(s−it,x)sσ−1ds. |
Clearly, we have the following estimate:
|˜p∗0(r−it,x)(σ)−˜p∗0,ε(r−it,x)(σ)|=1|Γ(σ)||∫ε0˜p(s−it,x)sσ−1ds|≲∫ε0s−d+Re(σ)−1ds≤1β−dεβ−d→ε→0+0uniformly for everyσ∈Sβ,γ, |
which proves that the sequence (˜p∗0,ε(r−it,x)(σ))ε>0 converges uniformly with limit ˜p∗0(r−it,x)(σ). Cauchy's theorem will guarantee that this limit is holomorphic.
Now, for s large, according to (3.1), we deduce that
|˜p(s−it,x)sσ−1|≲e−bss−D+Re(σ)−1, |
which ensures that the integral ∫∞1˜p(s−it,x)sσ−1ds converges uniformly on Re(σ)≤R, for each R∈R. In particular, it converges uniformly on compact subsets of C. It follows that the infinity part ˜p∗∞(r−it,x)(σ) represents an entire function of σ ([20, Ch. Ⅻ, Lemma 1.1, p. 308]).
Step 3. It remains to prove the meromorphic extension to the entire complex plane C of the function σ↦˜p∗(r−it,x)(σ)=w(t,x,σ,b) and to determine its poles. To do this, we proceed in the classical and usual way using an appropriate asymptotic expansion of the Poisson kernel.
To start, we state
p(s−it,x)∼cds−d∑j≥0αjsj,s→0+, | (3.5) |
where the coefficients αj=αj(t,x) are to be calculated later. Here, cd is a constant depending on d.
Given the Maclaurin series
e−bs∼∑j≥0(−1)jbjj!sj,s→0+, | (3.6) |
the asymptotic expansion of the product is given by
˜p(s−it,x)=e−bsp(s−it,x)∼cds−d∑j≥0˜αjsj,s→0+, | (3.7) |
where
˜αj=˜αj(t,x,b)=j∑l=0(−1)j−lbj−l(j−l)!αl(t,x),j∈N0. | (3.8) |
Now (3.7) is equivalent to asserting that for each nonnegative integer N, there exists a positive constant CN>0 such that
∀x∈X,∀t>0,∣˜p(s−it,x)−cdN∑j=0˜αjsj−d∣≤CNsN+1−d,0<s≤1. | (3.9) |
Then we decompose ˜p∗(r−it,x)(σ) for given σ with Re(σ)>d in the form
˜p∗(r−it,x)=w(t,x,σ,b)=AN(t,x,σ,b)+BN(t,x,σ,b)+˜p∗∞(r−it,x)(σ), | (3.10) |
where
● AN(t,x,σ,b)=1Γ(σ)∫10(˜p(s−it,x)−cdN∑j=0˜αjsj−d)sσ−1dt,
● BN(t,x,σ,b)=cdΓ(σ)N∑j=0˜αj∫10sj−d+σ−1ds=cdΓ(σ)N∑j=0˜αjσ+(j−d),
● ˜p∗∞(r−it,x)(σ) is an entire function of σ given by Eq (3.4).
Thus, the exact argument in [1,13] shows that AN(t,x,σ,b) is uniformly convergent on Re(σ)>d−(N+1)+ε for every ε>0. It follows that AN(t,x,σ,b) is a holomorphic function on Re(σ)>d−(N+1), and (3.10) provides the meromorphic continuation of ˜p∗(r−it,x)=w(t,x,σ,b) with simple poles at (σj=d−j)j∈N0 appearing in the function BN(t,x,σ,b), with residue given by
res(w(t,x,σ,b)),σj=d−j)=cdΓ(d−j)˜αj(t,x,b)=cdΓ(d−j)j∑l=0(−1)j−lbj−l(j−l)!αl(x,σ),∀x∈X=G/K, | (3.11) |
which ends the proof of Theorem 3.1.
Remark 3.2. Notice that, if j≥d, the integer d−j is negative; therefore, 1Γ(d−j)=0, and consequently res(w(t,x,σ,b),σj)=0, which allows us to affirm that w(t,x,σ,b) has a finite number of simple poles at the points σj=d−j, for j=0,1,⋯,(d−1), exactly as in the case of the ζ function on a compact manifold M [11].
Remark 3.3. Generally, we cannot compare the dimension d of X and D, the dimension at infinity of X without specifying the geometric structure of X. For example, if G complex, one has d=D, but when X has a normal real form, one has d<D, and in this case, the fundamental strip of ˜p∗(r−it,x)(σ)=w(t,x,σ,b) will be d<Re(σ)<D.
Let us turn to the computations of the coefficients ˜αj=˜αj(t,x,b). To do this, we refer to the key decomposition (3.10), which allows us to calculate the values of w(t,x,σ,b) for σ=−k,k∈N0 and to deduce an explicit expression of ˜αj, or even more simply an expression of αj.
Theorem 3.4. The coefficients (˜αj(t,x,b))t>0,x∈X are given by
˜αd+k(t,x,b)=c−1d(−1)kk!w(t,x,−k,b),k∈N0. | (3.12) |
Proof. For k∈N0 choose N>d+k in Eq. (3.10), which implies that −k>d−N>d−(N+1). It follows that AN(t,x,σ,b) is holomorphic at σ=−k, and since 1Γ(−k)=0, we obtain AN(t,x,−k,b)=0, and it is the same ˜p∗∞(r−it,x)(−k)=0. Now all that remains is to determine the value of BN(t,x,−k,b). Write
BN(t,x,−k,b)=cdΓ(−k)N∑j=0,j≠d+k˜αj−k+(j−d)⏟=0+f(σ)/σ=−k, | (3.13) |
where
f(σ)=cdΓ(σ)˜αd+kσ+k. | (3.14) |
Knowing that the only singularities of BN(t,x,σ,b) are at the points (σj=d−j)j=0,1,⋯,(d−1) and that
limσ→−kf(σ)=cd˜αd+kres(Γ(σ),−k)=cd˜αd+k(−1)kk!, |
we conclude that f is holomorphic at σ=−k and f(−k)=cd˜αd+k(−1)kk!. Eq (3.12) follows directly.
Remark 3.5. The results presented here are in agreement with general ζ-theory, but without any discussion on the parity of the space-dimension.
For the reader's benefit, we recall the definition of the singular expansion of a meromorphic function.
Definition 3.6. (Singular expansion) Let f(σ) be meromorphic in Q with P including all the poles of f(σ) in Q. A singular expansion of f(σ) in Q is a formal sum of singular elements of f(σ) at all points of P.
When S is a singular expansion of f(σ) in Q, we write
f(σ)≍S,σ∈Q. |
According to the previous paragraph and considering this last definition, it is easy to obtain the following result.
Theorem 3.7. The singular expansion of the wave kernel w(t,x,σ,b) is given by
w(t,x,σ,b)≍cdΓ(σ)d−1∑j=0˜αj(t,x,b)σ+(j−d),σ∈C,t>0,x∈X=G/K. | (3.15) |
Remark 3.8. It should be noted that the function BN(t,x,σ,b) in the decomposition (3.10) contains not only a singular part, but also a regular part. Indeed, since we chose N>d+k (therefore large enough) in the previous calculation, BN(t,x,σ,b) is written
BN(t,x,σ,b)=cdΓ(σ)d−1∑j=0˜αj(t,x,b)σ+(j−d)+cdΓ(σ)N∑j=d˜αj(t,x,b)σ+(j−d). |
Thus, a passage to the limit (N→∞) without worrying about convergence problems of the second sum, and by (3.12), we obtain the desired regular part
1Γ(σ)∑j≥0(−1)jj!w(t,x,−j,b)σ+j. |
As an application, we shall now see the analysis of harmonic sum assigned to the wave kernel on Riemannian symmetric spaces X=G/K. In particular, we derive a singular expansion of the Mellin transform of this sum, and we deduce an asymptotic expansion as s→0 of this sum using the converse correspondence rule (1.10). First, we introduce the general definition of harmonic sum, which will be involved in the asymptotic analysis of wave kernel. We refer to [21] for more details about sums of this type.
Definition 4.1. A sum of the form
G(s)=∑k≥1λkg(μks),s>0 | (4.1) |
is called a harmonic sum. The λk are the amplitudes, the μk are the frequencies, and g(s) is called the base function. The Dirichlet series of the harmonic sum is the sum given by
Λ(σ)=∑k≥1λkμ−σk,σ∈C. | (4.2) |
Here, we consider λk=1,μk=k and g(s)=˜p(s−it,x)=e−bsp(s−it,x) for s>0 and x∈X=G/K. In this case, Λ(σ) coincides with the zeta function ζ(σ)=∑k≥11kσ and
G(t,x,s,b)=∑k≥1˜p(ks−it,x)=∑k≥1e−bksp(ks−it,x). | (4.3) |
We start by calculating the Mellin transform of G(t,x,s,b).
Theorem 4.2. The Mellin transform of G(t,x,s,b) equals
G∗(t,x,σ,b)=ζ(σ)w(t,x,σ,b),t>0,x∈X=G/K, | (4.4) |
with fundamental strip (d,∞).
Proof. We apply the dominated convergence theorem two times. The first is dedicated to proving the convergence of G(t,x,s,b), while the second aims to obtain the separation formula (4.4).
Consider the partial sums
Gn(t,x,s,b)=n∑k=1˜p(ks−it,x)=n∑k=1e−bksp(ks−it,x),n≥1. |
A straightforward computation using a change of variables in the definition of G∗n(t,x,σ,b) and the inversion formula (1.3) gives
Gn(t,x,s,b)=∫a+i∞a−i∞Γ(σ)G∗n(t,x,σ,b)s−σdσ=∫a+i∞a−i∞Γ(σ)n∑k=11kσ˜p∗(r−it,x)(σ)s−σ⏟fn(σ)dσ,a>d. | (4.5) |
On the one hand, it is easy to see that
˜p∗(r−it,x)(σ)=O(1), |
since ˜p(s−it,x)=O(s−d) as s→0 and the integration is along the vertical line Re(σ)=a.
Then, this fact combined with the following well-known bounds
Γ(σ)∼√2π|Im(σ)|Re(σ)−1/2e−π2|Im(σ)|,Im(σ)→∞, | (4.6) |
and
|ζ(σ)|≲|Im(σ)|1−ε,ε<1, | (4.7) |
allow to dominate the sequence fn(σ) by an integrable function. More precisely, we obtain
|fn(σ)|≲|Im(σ)|1/2−ε+ae−π2|Im(σ)|s−a. | (4.8) |
Hence the convergence of G(t,x,s,b) by the dominated convergence theorem. Moreover, one has
G(t,x,s,b)=limn→∞Gn(t,x,s,b)=∫a+i∞a−i∞Γ(σ)ζ(σ)w(t,x,σ,b)s−σdσ. | (4.9) |
Now, it is clear by (4.8) that Gn(t,x,s,b)=O(s−a) and that G(t,x,s,b)=O(s−a), which makes it possible to apply the dominated convergence theorem to obtain via the definition of G∗ and interchange integral and limits
G∗(t,x,σ,b)=ζ(σ)w(t,x,σ,b),Re(σ)>d,x∈X=G/K, |
which ends the proof of the theorem.
Remark 4.3. A general proof of the previous theorem consists of taking into account a base function of fast decrease (see [21]), in order to use the dominated convergence theorem. Here we avoided this data and took advantage of Stirling's formula (4.6) to control the Gamma function present in the inversion formula (4.5).
Theorem 4.4. The asymptotic expansion of harmonic sum G(t,x,s,b) as s→0 is given by
G(t,x,s,b)∼1s(−cd˜αd−1(t,x,b)lns+a−1)+cdd−2∑j=0ζ(d−j)˜αj(t,x,b)sj−d, | (4.10) |
where a−1=res(Γ(σ)G∗(t,x,σ,b),σ=1).
Proof. First, notice that there are singularities at (σj=d−j)j=0,1,⋯,d−1. Globally, the poles of G∗(t,x,σ,b) are a double pole at σ=1 and simple at (σj=d−j)j=0,1,⋯,d−2.
According to the separation formula (4.4), we obtain the singular expansion
Γ(σ)G∗(t,x,σ,b)≍cd˜αd−1(t,x,b)(σ−1)2+a−1σ−1+cdd−2∑j=0ζ(d−j)˜αj(t,x,b)σ−(d−j). | (4.11) |
Again from the bound (4.8), we conclude that σrΓ(σ)G∗(t,x,σ,b)=O(1) as |σ|→∞, for all r>1. In other words, Γ(σ)G∗(t,x,σ,b) is of fast decrease.
Now all the ingredients are ready to be able to apply the correspondence rule A(σ−q)k+1↦(−1)kk!s−q(lns)k in singular expansion (4.11), and the asymptotic expansion is deduced directly.
Remark 4.5. It is not difficult to see that Γ(σ)G∗(t,x,σ,b) admits a meromorphic continuation to C and satisfies the conditions of the converse correspondence rule; thus, a complete asymptotic expansion of G(t,x,s,b) results.
Note that it is possible to express certain coefficients ˜αd+k(t,x,b) via the transform G∗(t,x,σ,b). More precisely, one has:
Theorem 4.6. The coefficients (˜αd+2k+1(t,x,b))t>0,x∈X are given by
˜αd+2k+1(t,x,b)=c−1d(2k+2)G∗(t,x,−2k−1,b)(2k+1)!B2k+2,k∈N0, | (4.12) |
where Bk denotes the Bernoulli numbers.
Proof. First, we know that
ζ(−k)=(−1)kBk+1k+1,k∈N0, | (4.13) |
and that
ζ(−2k)=0,B2k+1=0,k≥1, | (4.14) |
which implies that G∗(t,x,−2k,b)=0,∀k≥1.
For the odd negative integers, by (3.12) we write
w(t,x,−2k−1,b)=cd(−1)2k+1(2k+1)!˜αd+2k+1(t,x,b),k∈N0. |
Let us introduce this in the equation G∗(t,x,−2k−1,b)=ζ(−2k−1)w(t,x,−2k−1,b) well combined with (4.13), and the desired result is simply deduced.
Remark 4.7. Notice that Bernoulli numbers can appear in the singular expansion of Γ(σ)G∗(t,x,σ,b) given by formula (4.11). In fact, the function ϕ(s)=1es−1 admits an expansion near s=0
1es−1=∑k≥−1Bk+1(k+1)!sk, |
and by definition, its Mellin transform equals ζ(σ), which implies the singular expansion
Γ(σ)ζ(σ)≍∑k≥−1Bk+1(k+1)!1σ+k,σ∈C, |
leading to writing (4.11) differently, showing the numbers Bk, since one has Γ(σ)G∗(t,x,σ,b)=Γ(σ)ζ(σ)w(t,x,σ,b).
Remark 4.8. Consider the function
ψ(t,x,s,b)=∫∞01es/u−1˜p(u−it,x)duu,s>0,t>0,x∈X=G/K. |
It is easy to see that
ψ∗(σ)=Γ(σ)G∗(t,x,σ,b),Re(σ)>d. |
Thus, ψ(t,x,s,b) can be considered as the integral version of the harmonic series G(t,x,s,b). Consequently, it has the same asymptotic expansion (4.10), and its Mellin transform has the same singular expansion (4.11).
The results presented in this paper will be helpful in understanding the Mellin transform associated with the main evolution equations in Riemannian symmetric spaces of the non-compact type. We proved that the wave kernel σ↦w(t,x,σ,b)=˜p∗(r−it,x)(σ) extends meromorphically to the entire complex plane C with a finite number of simple poles on the real line, and we derived its singular expansion. In addition, we stated the singular expansion of the Mellin transform σ↦G∗(t,x,σ,b) of the harmonic sums G(t,x,s,b) to deduce its asymptotic expansion near s→0 by the converse correspondence rule.
In [1] we have studied the Mellin transform
w∗(z,x,σ,b)=1Γ(z)∫∞0e−btw(t,x,σ)tz−1dt, |
for z,σ∈C with Re(σ)>d, and a fixed element x∈X=G/K and a real parameter b>0, where w(t,x,σ) denotes the wave kernel on X=G/K defined by
w(t,x,σ)=1|W|∫a⋆eit√|λ|2+|ρ|2(|λ|2+|ρ|2)−σ/2φλ(x)|c(λ)|−2dλ, |
and we have obtained similar results as in subsection 3.1. In short, we can afford to configure the two studies, the previous and the current one, as follows:
˜w(t,x,σ)=e−btw(t,x,σ)→t→z˜w∗(z,x,σ,b), |
˜p(s−it,x)=e−bsp(s−it,x)→s→σ˜p∗(r−it,x)(σ)=w(t,x,σ,b). |
Thanks to our results furnished in Theorems 3.7, 4.2, and 4.4, we can elaborate on the same ones about singular expansion and harmonic sums. More precisely, if d=2p+1 is odd, we prove the following further results about the Mellin transform w∗(z,x,σ,b):
● Singular expansion:
˜w∗(z,x,σ,b)≍cdΓ(z)d−32∑j=0˜aj(x,σ,b)z+(j−d−12),Re(σ)>d,x∈X=G/K. |
● Harmonic sums:
H(t,x,σ,b)=∑k≥1e−bktw(kt,x,σ),x∈X=G/K, |
H∗(z,x,σ,b)=ζ(z)w∗(z,x,σ,b),Re(z)>d−12, |
Γ(z)H∗(z,x,σ,b)≍cd˜ad−32(x,σ,b)(z−1)2+a−1z−1+cdd−52∑j=0ζ(d−12−j)˜aj(x,σ,b)z−(d−12−j), |
H(t,x,σ,b)∼1t(−cd˜ad−32(x,σ,b)lnt+a−1)+cdd−52∑j=0ζ(d−12−j)˜aj(x,σ,b)tj−d−12,t→0, |
where a−1=res(Γ(z)H∗(z,x,σ,b),z=1).
We hope to return to all the questions proposed here in the special case when G is a complex semisimple Lie group, since the Harish-Chandra c-function and the spherical function φλ have elementary expressions, which can be a decisive factor to better express the coefficients αj(t,x,b).
The author declares he has not used Artificial Intelligence (AI) tools in the creation of this article.
The author extends his appreciation to Umm Al-Qura University, Saudi Arabia for funding this research work through grant number: 25UQU4361176GSSR01.
The author declares no conflict of interest.
[1] |
A. Hassani, On the meromorphic continuation of the Mellin transform associated to the wave kernel on Riemannian symmetric spaces of the non-compact type, AIMS Math., 9 (2024), 14731–14746. https://doi.org/10.3934/math.2024716 doi: 10.3934/math.2024716
![]() |
[2] | J. Bertrand, P. Bertrand, J. P. Ovarlez, The transforms and applications handbook, 2 Eds., CRC Press, 2000. |
[3] | Y. Brychkov, O. I. Marichev, N. V. Savischenko, Handbook of Mellin transforms, CRC Press, 2019. https://doi.org/10.1201/9780429434259 |
[4] | R. B. Paris, D. Kaminsky, Asymptotics and Mellin-Barnes integrals, Cambridge University Press, 2001. https://doi.org/10.1017/CBO9780511546662 |
[5] | V. A. Ditkin, A. P. Prudnikov, Integral transforms and operational calculus, New York: Pergamon Press, 1965. https://doi.org/10.1002/zamm.19660460819 |
[6] |
R. Cahn, J. Wolf, Zeta functions and their asymptotic expansions for compact symmetric spaces of rank one, Comment. Math. Helv., 51 (1976), 1–21. https://doi.org/10.1007/BF02568140 doi: 10.1007/BF02568140
![]() |
[7] |
R. Camporesi, On the analytic continuation of the Minakshisundaram-Pleijel zeta function for compact symmetric spaces of rank one, J. Math. Anal. Appl., 214 (1997), 524–549. https://doi.org/10.1006/jmaa.1997.5588 doi: 10.1006/jmaa.1997.5588
![]() |
[8] | T. F. Godoy, Minakshisundaram-Pleijel coefficients for compact locally symmetric spaces of classical type with non-positive sectional curvature, Ph.D. Thesis, Argentina: National University of Córdoba, 1987. |
[9] |
T. F. Godoy, R. J. Miatello, F. L. Williams, The local zeta function for symmetric spaces of non-compact type, J. Geom. Phys., 61 (2011), 125–136. https://doi.org/10.1016/j.geomphys.2010.08.008 doi: 10.1016/j.geomphys.2010.08.008
![]() |
[10] |
R. Miatello, On the Minakshisundaram-Pleijel coefficients for the vector-valued heat kernel on compact locally symmetric spaces of negative curvature, T. Am. Math. Soc., 260 (1980), 1–33. https://doi.org/10.2307/1999874 doi: 10.2307/1999874
![]() |
[11] |
S. Minakshisundaram, A. Pleijel, Some properties of the eigenfunctions of the Laplace operator on Riemannian manifolds, Can. J. Math., 1 (1949), 242–256. https://doi.org/10.4153/CJM-1949-021-5 doi: 10.4153/CJM-1949-021-5
![]() |
[12] |
F. L. Williams, Meromorphic continuation of Minakshisundaram-Pleijel series for semisimple Lie groups, Pac. J. Math., 182 (1998), 137–156. https://doi.org/10.2140/PJM.1998.182.137 doi: 10.2140/PJM.1998.182.137
![]() |
[13] |
F. L. Williams, Minakshisundaram-Pleijel coefficients for non-compact higher symmetric spaces, Anal. Math. Phys., 10 (2020). https://doi.org/10.1007/s13324-020-00396-x doi: 10.1007/s13324-020-00396-x
![]() |
[14] | M. Cowling, S. Guilini, S. Meda, Lp-Lq-Estimates for functions of the Laplace-Beltrami operator on noncompact symmetric spaces, Ⅱ, J. Lie Theory, 5 (1995), 1–14. |
[15] |
M. Cowling, S. Guilini, S. Meda, Lp-Lq-Estimates for functions of the Laplace-Beltrami operator on noncompact symmetric spaces, Ⅲ, Ann. I. Fourier, 51 (2001), 1047–1069. https://doi.org/10.5802/aif.1844 doi: 10.5802/aif.1844
![]() |
[16] |
M. Cowling, S. Guilini, S. Meda, Oscillatory multipliers related to the wave equation on noncompact symmetric spaces, J. Lond. Math. Soc., 66 (2002), 691–709. https://doi.org/10.1112/S0024610702003563 doi: 10.1112/S0024610702003563
![]() |
[17] | S. Helgason, Geometric analysis on symmetric spaces, American Mathematical Society, 1994. https://doi.org/10.1090/surv/039/02 |
[18] | S. Helgason, Groups and geometric analysis: Integral geometry, invariant differential operators, and spherical functions, American Mathematical Society, 2022. |
[19] | R. Gangolli, V. S. Varadarajan, Harmonic analysis of spherical functions on real reductive groups, Springer Science & Business Media, 2012. https://doi.org/10.1007/978-3-642-72956-0 |
[20] | S. Lang, Complex analysis, 2 Eds., Springer Verlag, 1985. https://doi.org/10.1007/978-1-4757-1871-3 |
[21] |
P. Flajolet, X. Gourdon, P. Dumas, Mellin transform and asymptotics: Harmonic sums, Theor. Comput. Sci., 144 (1995), 3–58. https://doi.org/10.1016/0304-3975(95)00002-E doi: 10.1016/0304-3975(95)00002-E
![]() |