Home Mathematics On an evolution equation in sub-Finsler geometry
Article Open Access

On an evolution equation in sub-Finsler geometry

  • Nicola Garofalo EMAIL logo
Published/Copyright: November 11, 2024

Abstract

We study the gradient flow of an energy with mixed homogeneity, which is at the interface of Finsler and sub-Riemannian geometry.

MSC 2010: 35H20; 35B09; 35R03; 53C17; 58J60

1 Introduction

Singular spaces occupy a prominent position in analysis and geometry. Examples of basic interest are Alexandrov spaces and Finsler manifolds. These latter ambients have received considerable attention over the past few decades as they often occur as scaling limits of crystalline or Riemannian structures, see [21] and especially the seminal work [3]. In this article, we introduce a nonlinear evolution equation that represents the gradient flow of an energy which is at the interface of Finsler and sub-Riemannian geometry. Our main objective is to understand in detail the relevant heat kernel. To keep the presentation self-contained, we confine ourselves to the model setting of a product of Euclidean spaces, i.e., R N = R m × R k , with coordinates z R m , σ R k . We assume that on each of the two layers, R m and R k , Minkowski norms Φ and Ψ have been assigned. By this, we mean that Φ 2 C 2 ( R m \ { 0 } ) , Ψ 2 C 2 ( R k \ { 0 } ) , and that these functions are strongly convex and 1-homogeneous. We, respectively denote by Φ 0 and Ψ 0 their dual norms, defined as in (2.2), and acting on the dual spaces of R m and R k , which are in turn identified with R m and R k themselves, with coordinates z and σ , respectively. The level sets of the functions Φ 0 and Ψ 0 are often referred to as Wulff shapes. This situation represents a simplified, yet significant, model for the more general one of a Riemannian manifold in which the tangent space is stratified in layers, each endowed with a different anisotropic structure, and having mixed homogeneities weighted according to the relative position in the stratification.

We are interested in the L 2 gradient flow

(1.1) f t = E Φ , Ψ ( f ) f

of the following energy:

(1.2) E Φ , Ψ ( f ) = 1 2 R N Φ ( z f ) 2 + Φ 0 ( z ) 2 4 Ψ ( σ f ) 2 d z d σ .

Notable features of (1.2) are as follows:

  1. The nonlinear dependence on the degenerate gradient

    (1.3) X f = ( X 1 f , , X m f , X m + 1 f , , X N f )

    associated with the N vector fields

    (1.4) X i = z i , i = 1 , , m , X m + j = Φ 0 ( z ) 2 σ j , j = 1 , , k

    (note from (1.4) that for i = 1 , , N one has X i = X i in L 2 ( R N ) . We also note that, under the given assumptions on Φ , we have Φ 0 C 1 ( R m \ { 0 } ) , see [20, Corollary 1.7.3].

  2. It degenerates along the manifold M = { 0 } z × R k , but because of the anisotropic nature of the dual norm Φ 0 ( z ) , it does so at different scales along regions of approach to M .

  3. It is invariant with respect to the family of mixed dilations in R N

    (1.5) δ λ ( z , σ ) = ( λ z , λ 2 σ ) , λ > 0 .

Returning to (1.1), a standard argument shows that the relevant evolution partial differential equation (PDE) attached to (1.2) is

(1.6) f t = Δ Φ ( f ) + Φ 0 ( z ) 2 4 Δ Ψ ( f ) ,

where we have denoted by Δ Φ and Δ Ψ the Finsler Laplacians in the spaces R m and R k , see (2.6). Furthermore, if with δ λ as in (1.5), we define

(1.7) D λ ( z , σ , t ) = ( δ λ ( z , σ ) , λ 2 t ) ,

then if f solves (1.6), so does also f D λ , for every λ > 0 .

We emphasise that (1.6) is a nonlinear evolution equation with quadratic growth in the degenerate “gradient” (1.3). By this, we mean that there exist constants γ , γ > 0 such that

(1.8) γ X f 2 A ( X f ) , X f γ X f 2 .

To explain (1.8), note that (1.3) and (1.4) give

(1.9) X f = z f Φ 0 ( z ) 2 σ f .

Therefore, thanks to the 1-homogeneity of Φ and Ψ , if we let

A ( X f ) = Φ ( z f ) Φ ( z f ) Φ 0 ( z ) 2 Ψ ( σ f ) Ψ ( η ) ,

then the PDE (1.6) can be alternatively written as

(1.10) f t = i = 1 N X i A i ( X f ) ,

where the functions A i : R N R are the components of the vector field

A ( X f ) = ( A 1 ( X f ) , , A N ( X f ) ) T .

We now note that

(1.11) A ( X f ) , X f = Φ ( z f ) Φ ( z f ) , z f + Φ 0 ( z ) 2 4 Ψ ( σ f ) Ψ ( σ f ) , σ f = Φ ( z f ) 2 + Φ 0 ( z ) 2 4 Ψ ( σ f ) 2 ,

where in the last equality, we have used the 1-homogeneity of Φ and Ψ , which is well-known to be equivalent to the Euler equations

ξ , Φ ( ξ ) = Φ ( ξ ) , η , Ψ ( η ) = Ψ ( η ) .

By the condition (2.7), from (1.9) and (1.11), we finally infer the existence of γ , γ > 0 such that (1.8) holds. Although the present work is not concerned with the local theory of the PDE (1.6), we remark that the growth (1.8) would allow us to apply the results in [4] once the relevant volume doubling condition and Poincaré inequality are available. For these aspects, see the seminal works [13,19], and also the more recent article [5] that develops the local theory in Finsler manifolds with Ricci lower bounds.

To state our first result, we need to introduce some notations. Henceforth, points of R m will be denoted by z , ζ , etc., points of R k by σ , τ , λ , etc. We will use the notation X = ( z , σ ) , Y = ( ζ , τ ) , etc., for points in the product space R N . Since the notation X will no longer appear in this work, there will be no risk of confusion of this notation with the subscript in (1.8), (1.9), and (1.11). Also, we will, respectively, indicate with σ Φ and σ Ψ the intrinsic measures of the Wulff spheres defined by

(1.12) σ Φ = { Φ 0 ( z ) = 1 } d H m 1 ( z ) Φ 0 ( z ) , σ Ψ = { Ψ 0 ( σ ) = 1 } d H k 1 ( σ ) Ψ 0 ( σ ) ,

where H m 1 and H k 1 denote ( m 1 ) -dimensional and ( k 1 ) -dimensional Hausdorff measure in R m and R k , respectively. Recall the classical formulas σ m 1 = H m 1 ( S m 1 ) = 2 π m 2 Γ m 2 , σ k 1 = H k 1 ( S k 1 ) = 2 π k 2 Γ k 2 . It is clear from (1.12) that, in the linear isotropic case Φ ( z ) = z , Ψ ( σ ) = σ , one has σ Φ = σ m 1 , σ Ψ = σ k 1 . For a number ν C , we will denote with J ν the Bessel function of the first kind and order ν and indicate by

(1.13) G ν ( z ) = z ν J ν ( z ) .

We have the following.

Theorem 1.1

For every X = ( z , σ ) R N and t > 0 , the function

(1.14) G ( X , t ) = σ m 1 σ k 1 σ Φ σ Ψ ( 2 π ) k 2 ( 4 π ) m 2 t m 2 + k 0 u sinh u m 2 e u tanh u Φ 0 ( z ) 2 4 t G k 2 1 u Ψ 0 ( σ ) t u k 1 d u

is a solution of equation (1.6). Moreover, for every t > 0 , we have

(1.15) R N G ( X , t ) d X = 1 .

Remark 1.2

We emphasise that Theorem 1.3 proves that, in fact, function (1.14) is a fundamental solution for (1.6).

To state our second result, we mention that in recent paper [7], the authors have studied in the product space R N the following degenerate energy with mixed homogeneity:

(1.16) E α , p ( u ) = 1 p R N Φ ( z u ) 2 + Φ 0 ( z ) 2 α 4 Ψ ( σ u ) 2 p 2 d z d σ , 1 < p < .

It is clear that (1.2) corresponds to the special case α = 1 and p = 2 of (1.16). One of their main results is an explicit fundamental solution for the Euler-Lagrange equation of (1.16). To formulate the relevant result, consider the following anisotropic Minkowski gauge:

(1.17) Θ ( z , σ ) = ( Φ ( z ) 2 ( α + 1 ) + 4 ( α + 1 ) 2 Ψ ( σ ) 2 ) 1 2 ( α + 1 ) .

In [7], a new Legendre transformation Θ 0 was introduced, namely

(1.18) Θ 0 ( z , σ ) α + 1 = sup Θ ( ξ , τ ) = 1 ( z , ξ α + 1 + 4 ( α + 1 ) 2 σ , τ ) .

The remarkable feature of (1.18) is underscored by the following result, established in [7, Proposition 3.3]:

(1.19) Θ 0 ( z , σ ) = ( Φ 0 ( z ) 2 ( α + 1 ) + 4 ( α + 1 ) 2 Ψ 0 ( σ ) 2 ) 1 2 ( α + 1 ) ,

where Φ 0 and Ψ 0 are the classical dual Minkowski norms defined as in (2.2). The reader should observe the perfect symmetry between (1.17) and (1.19). Furthermore, in [7, Theorem 1.2], the authors proved that the function

(1.20) E α , p ( z , σ ) = C α , p Θ 0 ( z , σ ) Q p p 1 , p Q , C α log Θ 0 ( z , σ ) , p = Q

is a fundamental solution, with pole in ( 0 , 0 ) , of the Euler-Lagrange equation of (1.16). The number Q in (1.20) is given by

(1.21) Q = Q α = m + ( α + 1 ) k .

It is clear from (1.20) that such number plays the role of a dimension. In (1.20), the positive constants C α , p and C α are implicitly given, and they involve the Wulff shapes of the gauge Θ 0 . However, in order to fully understand the gradient flow of the energy (1.2), it is important to have an explicit knowledge of the constants. For this reason, in Section 3, we turn to this problem, and in Lemma 3.4, we show that, with σ Φ and σ Ψ as in (1.12), then

(1.22) σ α , p = σ Φ σ Ψ B m + α p 2 ( α + 1 ) , k 2 2 k + 1 ( α + 1 ) k ,

where we have denoted by B ( x , y ) Euler beta function (3.7). Since in [7, Theorem 1.2] it was proved that

(1.23) C α , p = p 1 Q p ( σ α , p ) 1 ( p 1 ) , p Q , σ α , Q 1 ( Q 1 ) , p = Q ,

it is clear that (1.22) provides an expression of C α , p . We explicitly note from (1.19), (1.20), and (1.21) that, when α = 1 and p = 2 , we have

(1.24) Θ 0 ( z , σ ) = ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 ) 1 4

and

(1.25) Q = m + 2 k .

We also mention that the above theorem (1.20) generalised a result first established in [10] for the linear case Φ ( z ) = z , Ψ ( σ ) = σ , when p = 2 .

To put our second result in some historical context, consider the standard Gauss-Weierstrass kernel in R n (here, we are assuming n 2 )

G ( x , t ) = ( 4 π t ) n 2 e x 2 4 t , t > 0 .

A well-known manifestation of the fact that G ( x , t ) is a fundamental solution of the heat equation, and of the Bochner subordination principle, is that the L loc 1 ( R N ) function defined by

E ( x ) = def 0 G ( x , t ) d t

provides a fundamental solution of Δ with pole in 0 R n . Furthermore, an elementary, yet beautiful computation shows that

E ( x ) = 1 ( n 2 ) σ n 1 x 2 n .

The next theorem shows that, surprisingly, despite its strongly nonlinear character, the evolution equation (1.6) displays the same linear phenomenon as the standard heat equation. Furthermore, Theorem 1.3 shows the remarkable fact that, had we not known the magic Minkowski gauge Θ 0 in (1.24), by running the nonlinear heat flow (1.1), we are forced to discover it!

Theorem 1.3

Let G ( X , t ) be the function defined in (1.14) of Theorem 1.1. For every t > 0 , we have

(1.26) 0 G ( X , t ) d t = C 1 , 2 Θ 0 ( z , σ ) Q 2 ,

where Θ 0 is as in (1.24), Q as in (1.25), and C 1 , 2 is the constant, identified by (1.22), and corresponding to the case α = 1 and p = 2 in (1.23). As a consequence of (1.20), the function defined by the left-hand side of (1.26) is a fundamental solution of the nonlinear equation

(1.27) Δ Φ ( u ) + Φ 0 ( z ) 2 4 Δ Ψ ( u ) = 0 .

This proves that G ( X , t ) is not only a solution to (1.6), but in fact it is a fundamental solution.

We close this introduction with a short description of the organisation of the article. In Section 2, we collect some basic properties of Minkowski norms and of the Finsler Laplacian and heat equation. Section 3 is devoted to proving Proposition 3.5. In Section 4, we prove Theorem 1.1. Finally, in Section 5, we establish Theorem 1.3.

2 Minkowski norms

Let M : R n [ 0 , ) be a Minkowski norm in R n . By this, we mean that M 2 C 2 ( R n \ { 0 } ) is a strongly convex function such that M ( λ x ) = λ M ( x ) for every x R n and λ R . By strong convexity, we mean that the Hessian matrix 1 2 2 ( M 2 ) is positive definite in R n \ { 0 } . Since all norms in R n are equivalent, there exist constants β α > 0 such that

(2.1) α ξ M ( ξ ) β ξ .

We denote by

(2.2) M 0 ( x ) = sup M ( ξ ) = 1 x , ξ

its Legendre transform, also known as the dual norm of M . The Cauchy-Schwarz inequality trivially holds

x , y M ( x ) M 0 ( y ) .

A basic property of the norms M and M 0 is the following, see [3, Lemma 2.1] and also (3.12) in [6],

(2.3) M ( M 0 ( x ) ) = M 0 ( M ( x ) ) = 1 , x R n \ { 0 } .

Given a function u C 1 ( R n ) , an elementary, yet useful, consequence of the homogeneity of M is

(2.4) M ( u ( x ) ) , u ( x ) = M ( u ( x ) ) .

A less obvious basic fact is the following formula, which can be found in [3, Lemma 2.2].

Lemma 2.1

For every x R n \ { 0 } , one has

M 0 ( x ) M ( M 0 ( x ) ) = x , M ( x ) M 0 ( M ( x ) ) = x .

The Euler-Lagrange equation of the energy

(2.5) E M ( u ) = 1 2 M ( u ) 2 d x

is the so-called Finsler Laplacian

(2.6) Δ M ( u ) = div ( M ( u ) M ( u ) ) = 0 .

It is worth emphasising here that the operator in (2.6) is quasilinear, but not linear, unless of course M ( x ) = x . However, the operator Δ M is elliptic. In fact, since M 2 is homogeneous of degree 2, the Euler formula gives

( M 2 ) ( ξ ) , ξ = M ( ξ ) 2 ,

and from (2.1), we thus have for every ξ R n

(2.7) α 2 ξ 2 ( M 2 ) ( ξ ) , ξ β 2 ξ 2 .

An important property of the dual norm is the following result that follows from (2.3) and from Lemma 2.1, see [6] and [9].

Proposition 2.2

Consider the function

ψ ( x ) = M 0 ( x ) .

Then, if k C 2 ( R ) , and v = k ψ , one has in R n \ { 0 }

Δ M ( v ) = k ( ψ ) + n 1 ψ k ( ψ ) .

A remarkable consequence of Proposition 2.2 is that the nonlinear operator Δ M acts linearly on functions of the dual norm M 0 . The Finsler heat equation arises from the gradient flow of the energy (2.5). It is the quasilinear PDE in R n × R

(2.8) t u = Δ M ( u ) .

In the framework of Finsler manifolds, an in-depth study of (2.8) has been carried out in the previous studies [17,18], see also [1] for the case of R n . Using Proposition 2.2, it is possible to construct the following notable explicit solution of the Finsler heat equation, see [17, Example 4.3] and also [1].

Proposition 2.3

The function

G ( x , t ) = t n 2 e M 0 ( x ) 2 4 t , x R n , t > 0 ,

solves the heat equation in R n × ( 0 , ) , i.e.,

t G Δ M ( G ) = 0 .

We leave it as an exercise for the reader to verify that the function G ( x , t ) satisfies Proposition 2.4, the Finsler counterpart of the extremal case of the famous result in [15]. A Li-Yau theory in Finsler manifolds was developed in the cited work [18]. In connection with the present work, a version of the theory in [2] will be presented in a forthcoming article.

Proposition 2.4

The following identity is true in R n × ( 0 , )

M ( log G ) 2 t ( log G ) = n 2 t .

Using the coarea formula, one can easily show that

(2.9) R n G ( x , t ) d x = 2 n 1 Γ ( n 2 ) σ M ,

where we have let

(2.10) σ M = { M 0 ( x ) = 1 } d H n 1 ( x ) M 0 ( x ) .

Note that, on the one hand, the coarea formula gives

(2.11) Vol n ( { x R n Φ 0 ( x ) < r } ) = 0 r { M 0 ( x ) = s } d H n 1 ( x ) M 0 ( x ) d s ,

and therefore,

(2.12) d d r Vol n ( { x R n Φ 0 ( x ) < r } ) = { M 0 ( x ) = r } d H n 1 ( x ) M 0 ( x ) .

On the other hand, a rescaling gives

(2.13) Vol n ( { x R n Φ 0 ( x ) < r } ) = ω M r n ,

and therefore,

(2.14) d d r Vol n ( { x R n Φ 0 ( x ) < r } ) = n ω M r n 1 .

Equating (2.12) and (2.14), we infer

(2.15) { M 0 ( x ) = r } d H n 1 ( x ) M 0 ( x ) = n ω M r n 1 , r > 0 .

In particular, when r = 1 , we obtain from (2.10) and (2.15)

(2.16) σ M = n ω M .

Remark 2.5

The identity (2.16) can also be directly obtained by the well-known formula of Minkowski for the volume, which gives

(2.17) ω M = 1 n { M 0 ( x ) = 1 } x , ν d H n 1 ( x ) .

If we now use the first identity in Lemma 2.1, the fact that ν = M 0 M 0 , and the Euler formula M ( p ) , p = M ( p ) , applied with p = M 0 ( x ) , we find on { M 0 ( x ) = 1 }

x , ν = M 0 ( x ) M ( M 0 ( x ) ) , M 0 ( x ) M 0 ( x ) = M ( M 0 ( x ) ) M 0 = 1 M 0 ,

where in the last equality we have used (2.3). Substituting in (2.17), and using (2.10), we obtain (2.16).

We will need the following useful observation.

Lemma 2.6

Suppose f ( x ) = f ( x ) , for some measurable function f : [ 0 , ) R , and consider the function F ( x ) = f ( M 0 ( x ) ) . Then,

R n F ( x ) d x = σ M σ n 1 R n f ( x ) d x ,

where as customary σ n 1 = 2 π n 2 Γ ( n 2 ) .

Proof

The coarea formula and (2.15) give

R n F ( x ) d x = 0 f ( r ) { M 0 ( x ) = r } d H n 1 ( x ) M 0 ( x ) d r = σ M 0 f ( r ) r n 1 d r = σ M σ n 1 R n f ( x ) d x .

3 Equalities matter

This section is devoted to the explicit computation of constants C α , p and C α in the above cited (1.20) from [7, Theorem 1.2]. The main result of the section is Proposition 3.5. This result plays a critical role in recognising that the function G ( X , t ) introduced in (1.14) of Theorem 1.1 is, in fact, a fundamental solution of the PDE (1.6). Throughout this section, R N = R m × R k , and Φ and Ψ will be strongly convex Minkowski norms in R m and R k , respectively. We recall from [7] that, with Θ 0 ( z , σ ) given by (1.19), the constants in (1.20) are prescribed by the formulas (1.23), where, with Q as in (1.21), we have let

(3.1) σ α , p = Q { Θ 0 ( z , σ ) < 1 } Φ 0 ( z ) Θ 0 ( z , σ ) α p d z d σ .

If we now consider the Wulff ball of the anisotropic gauge

K = { ( z , σ ) R N Θ 0 ( z , σ ) < 1 }

and set

K = { ( r , s ) [ 0 , ) × [ 0 , ) r 2 ( α + 1 ) + 4 ( α + 1 ) 2 s 2 < 1 } ,

then it should be clear to the reader that, if we define

(3.2) f ( r , s ) = r α p ( r 2 ( α + 1 ) + 4 ( α + 1 ) 2 s 2 ) α p 2 ( α + 1 ) 1 K ( r , s ) ,

where we have denoted by 1 K the indicator function of the set K , then we have

(3.3) F ( z , σ ) = Φ 0 ( z ) Θ 0 ( z , σ ) α p 1 K ( z , σ ) = f ( Φ 0 ( z ) , Ψ 0 ( σ ) ) .

We now state a generalisation of Lemma 2.6, whose proof we leave to the reader.

Lemma 3.1

Let f : [ 0 , ) × [ 0 , ) R ¯ be a measurable function, and let

F ( z , σ ) = f ( Φ 0 ( z ) , Ψ 0 ( σ ) ) , f ( z , σ ) = f ( z , σ ) .

Then,

(3.4) R N F ( z , σ ) d z d σ = σ Φ σ Ψ σ m 1 σ k 1 R N f ( z , σ ) d z d σ .

Furthermore, one has

(3.5) R N f ( z , σ ) d z d σ = σ m 1 σ k 1 0 0 f ( r , s ) r m 1 s k 1 d r d s .

Therefore, by combining (3.4) and (3.5), we find

(3.6) R N F ( z , σ ) d z d σ = σ Φ σ Ψ 0 0 f ( r , s ) r m 1 s k 1 d r d s .

If we now apply (3.6), keeping (3.1), (3.2), and (3.3) in mind, we reach the following conclusion.

Lemma 3.2

For every α , m , k , > 0 , with Q given as in (1.21), we have

σ α , p = σ Φ σ Ψ Q K r α p ( r 2 ( α + 1 ) + 4 ( α + 1 ) 2 s 2 ) α p 2 ( α + 1 ) r m 1 s k 1 d r d s .

In the next lemma, we compute in closed form the integral on the right-hand side of the equation in Lemma 3.2. Such a result will be crucial in the proof of Proposition 3.5. In its statement, we denote by

(3.7) B ( x , y ) = 2 0 π 2 ( cos ϑ ) 2 x 1 ( sin ϑ ) 2 y 1 d ϑ

Euler beta function. For the reader’s convenience, we recall its well-known property

(3.8) B ( x , y ) = Γ ( x ) Γ ( y ) Γ ( x + y ) .

Lemma 3.3

For any p > 1 , and m , k > 0 , one has

K r α p ( r 2 ( α + 1 ) + 4 ( α + 1 ) 2 s 2 ) α p 2 ( α + 1 ) r m 1 s k 1 d r d s = B m + α p 2 ( α + 1 ) , k 2 2 k + 1 ( α + 1 ) k ( m + ( α + 1 ) k ) .

Proof

Consider the open quadrant Ω = ( 0 , ) × ( 0 , ) in the ( r , s ) -plane. Define a mapping from the open strip U = ( 0 , ) × ( 0 , π 2 ) of the ( ρ , ϑ ) -plane onto Ω by the equations

r = ρ ( cos ϑ ) 1 α + 1 , s = ρ α + 1 2 ( α + 1 ) sin ϑ .

The Jacobian determinant of such mapping is

J ( ρ , ϑ ) = ρ α + 1 2 ( α + 1 ) ( cos ϑ ) α α + 1 ,

therefore ( ρ , ϑ ) ( r , s ) is a diffeomorphism of U onto Ω . The Jacobi formula for the change of variable and the defining equation (3.7) thus give

K r α p ( r 2 ( α + 1 ) + 4 ( α + 1 ) 2 s 2 ) α p 2 ( α + 1 ) r m 1 s k 1 d r d s = 1 2 k ( α + 1 ) k 0 1 ρ m + ( α + 1 ) k 1 d ρ 0 π 2 ( cos ϑ ) m + α p α + 1 1 ( sin ϑ ) k 1 d ϑ = 1 2 k + 1 ( α + 1 ) k B ( x , y ) m + ( α + 1 ) k ,

where

2 x 1 = m + α p α + 1 1 , 2 y 1 = k 1 .

This completes the proof.□

Combining Lemmas 3.2 and 3.3, we obtain the following result.

Lemma 3.4

For every m , k N , α > 0 , and p > 1 , one has

(3.9) σ α , p = σ Φ σ Ψ B m + α p 2 ( α + 1 ) , k 2 2 k + 1 ( α + 1 ) k .

In connection with Theorem 1.3, we are particularly interested in the case α = 1 and p = 2 of Lemma 3.4. Our objective is to establish the following important fact.

Proposition 3.5

When α = 1 and p = 2 , the constant in (1.23) admits the following alternate representation:

(3.10) C 1 , 2 = 1 ( Q 2 ) σ 1 , 2 = σ m 1 σ k 1 σ Φ σ Ψ 2 m 2 + 2 k 4 Γ m 4 Γ 1 2 m 2 + k 1 π m + k + 1 2 .

As a consequence of (3.10), and of the above cited (1.20) from [7], the function defined by the integral in the left-hand side of (1.26) is the fundamental solution of the nonlinear PDE (1.27).

Proof

It is clear that, to prove the proposition, it suffices to show that

(3.11) σ 1 , 2 = σ Φ σ Ψ σ m 1 σ k 1 π m + k + 1 2 ( m + 2 k 2 ) 2 m 2 + 2 k 4 Γ m 4 Γ 1 2 m 2 + k 1 ,

where we have used (1.25) to write Q 2 = m + 2 k 2 . Since according to (3.9), we have

σ 1 , 2 = σ Φ σ Ψ B m + 2 4 , k 2 2 2 k + 1 ,

keeping (3.8) in mind, we infer that, in order to establish (3.11), we need to prove the following identity:

(3.12) Γ m 4 + 1 2 Γ k 2 Γ m + 2 k + 2 4 = 1 σ m 1 σ k 1 π m + k + 1 2 ( m + 2 k 2 ) 2 m 2 5 Γ m 4 Γ 1 2 m 2 + k 1 .

Since σ m 1 = 2 π m 2 Γ m 2 , σ k 1 = 2 π k 2 Γ k 2 , we see that (3.12) is equivalent to proving that

(3.13) Γ 1 2 m 2 + 1 Γ 1 2 m 2 + k + 1 = π Γ m 2 ( m + 2 k 2 ) 2 m 2 3 Γ m 4 Γ 1 2 m 2 + k 1 .

At this point, we use twice the Legendre duplication formula (see, e.g. [14, (1.2.3) on p.3])

(3.14) 2 2 x 1 Γ ( x ) Γ x + 1 2 = π Γ ( 2 x ) .

The first time, we take x = 1 2 m 2 + 1 , for which we have

2 x 1 = m 2 , x + 1 2 = m 4 + 1 ,

obtaining

(3.15) Γ 1 2 m 2 + 1 = π Γ m 2 2 m 2 1 Γ m 4 .

For the reader’s convenience, we mention that we have repeatedly used the well-known formula Γ ( ν + 1 ) = ν Γ ( ν ) . The second time, we take x = 1 2 m 2 + k + 1 , which gives

2 x 1 = m 2 + k , x + 1 2 = 1 2 m 2 + k + 1 .

We thus find

(3.16) 1 Γ 1 2 m 2 + k + 1 = 2 m 2 + k 1 Γ 1 2 m 2 + k π Γ m 2 + k .

Combining (3.15) with (3.16), we see that now (3.13) is reduced to verifying whether the identity

(3.17) 2 m 2 + k 3 Γ 1 2 m 2 + k 1 Γ 1 2 m 2 + k = π Γ m 2 + k ( m + 2 k 2 )

is true or not. This can be accomplished by one last application of (3.14), this time with x = 1 2 m 2 + k 1 , for which

2 x 1 = m 2 + k 2 , x + 1 2 = 1 2 m 2 + k .

We thus find that (3.17) becomes equivalent to verifying that the following identity holds:

Γ m 2 + k 1 2 = Γ m 2 + k m + 2 k 2 .

This is obviously true as a consequence of Γ ( ν + 1 ) = ν Γ ( ν ) . We have thus completed the proof of Proposition 3.5.□

4 Construction of the heat kernel

In this section, we prove Theorem 1.1. We begin by recalling that, with G ν defined by (1.13), if h : R + C is a measurable function, its Hankel transform of order ν is defined by

(4.1) ν ( h ) ( s ) = ( 2 π ) ν + 1 0 h ( u ) G ν ( 2 π u s ) u 2 ν + 1 d u ,

see p. 18 in [16], and also [11, Sec. 22]. Suppose now that h ( λ ) = h ( λ ) , with λ R k is a spherically symmetric function in R k . Its Fourier transform is given by the well-known Bochner formula

(4.2) h ˆ ( σ ) = 2 π σ k 2 1 0 h ( u ) J k 2 1 ( 2 π σ u ) u k 2 d u .

Comparing (4.2) with (4.1), we see that the Hankel representation of h ˆ is given by

(4.3) h ˆ ( σ ) = k 2 2 ( h * ) ( σ ) = ( 2 π ) k 2 0 h ( u ) G k 2 1 ( 2 π u σ ) u k 1 d u .

We next recall the following result, see [12, Theorem 3.4].

Theorem 4.1

Given the PDE in R N × R ,

(4.4) t f = Δ z f + z 2 4 Δ σ f ,

its heat kernel with pole at X = ( z , σ ) R N is given by

H ( X , X , t ) = 2 k ( 4 π t ) m 2 + k R k e i t λ , σ σ λ sinh λ m 2 e λ 4 t tanh λ ( z 2 + z 2 2 z , z sech λ ) d λ .

The fact that H ( X , X , t ) is the heat kernel of (4.4) means that, for every fixed X R N , the function ( X , t ) H ( X , X , t ) solves (4.4) in R N × ( 0 , ) , and that furthermore one has

R N H ( X , X , t ) φ ( X ) d X t 0 + φ ( X ) .

Henceforth, when X = 0 R N , we use the abbreviated notation H ( X , t ) = H ( X , 0 , t ) , so that

(4.5) H ( X , t ) = 2 k ( 4 π t ) m 2 + k R k e i t σ , λ λ sinh λ m 2 e λ tanh λ z 2 4 t d λ , t > 0 .

It is clear from (4.3) and (4.5) that, if we consider in R k the rapidly decreasing spherically symmetric function

(4.6) λ h ( λ ) = h ( λ ) = λ sinh λ m 2 e λ tanh λ z 2 4 t ,

then

(4.7) H ( X , t ) = 2 k ( 4 π t ) m 2 + k h ˆ σ 2 π t ,

and therefore we have the following.

Lemma 4.2

The function in (4.5) can be written as

(4.8) H ( X , t ) = ( 2 π ) k 2 ( 4 π ) m 2 t m 2 + k 0 u sinh u m 2 e u tanh u z 2 4 t G k 2 1 u σ t u k 1 d u .

We reiterate that the function H ( X , t ) solves in R N × ( 0 , ) the PDE (4.4), i.e.,

(4.9) H t = Δ z H + z 2 4 Δ σ H .

We can now provide the proof of Theorem 1.1.

Proof of Theorem 1.1

It is clear that H ( X , t ) is spherically symmetric in both z R m and σ R k . In fact, from (4.8) in Lemma 4.2, we can write it in the form

(4.10) H ( X , t ) = F ( z , σ , t ) ,

where

(4.11) F ( r , s , t ) = ( 2 π ) k 2 ( 4 π ) m 2 t m 2 + k 0 u sinh u m 2 e u tanh u r 2 4 t G k 2 1 u s t u k 1 d u .

From the fact that H ( X , t ) solves the PDE (4.9) in R N × ( 0 , ) , we infer that the function (5.6) satisfies the following PDE in the variables ( r , s , t ) R + × R + × ( 0 , )

(4.12) F t = F r r + m 1 r F r + r 2 4 F s s + k 1 s F s .

Since, on the other hand, it is obvious from (1.14) and (4.11) that

(4.13) G ( X , t ) = σ m 1 σ k 1 σ Φ σ Ψ F ( Φ 0 ( z ) , Ψ 0 ( σ ) , t ) ,

from the chain rule and a double application of Proposition 2.2, we infer

(4.14) Δ Φ ( G ) + Φ 0 ( z ) 2 4 Δ Ψ ( G ) ( X , t ) = σ m 1 σ k 1 σ Φ σ Ψ F r r + m 1 r F r + r 2 4 F s s + k 1 s F s ( Φ 0 ( z ) , Ψ 0 ( σ ) , t ) = σ m 1 σ k 1 σ Φ σ Ψ F t ( Φ 0 ( z ) , Ψ 0 ( σ ) , t ) = G t ( X , t ) ,

where in the second to the last equality we have used (4.12). Equation (4.14) shows that the function G ( X , t ) solves the nonlinear PDE (1.6).

We next prove (1.15). We have from (1.14) and Fubini’s theorem

R N G ( X , t ) d X = σ m 1 σ k 1 σ Φ σ Ψ ( 2 π ) k 2 ( 4 π ) m 2 t m 2 + k R k 0 u sinh u m 2 G k 2 1 u Ψ 0 ( σ ) t u k 1 × R m e u tanh u Φ 0 ( z ) 2 4 t d z d u d σ .

By the homogeneity of Φ 0 and a change of variable, we have

R m e u tanh u Φ 0 ( z ) 2 4 t d z = R m e Φ 0 ( u tanh u 1 4 t z ) 2 d z = tanh u u m 2 2 m t m 2 R m e Φ 0 ( z ) 2 d z .

Now, note that Lemma 2.6 gives

R m e Φ 0 ( z ) 2 d z = σ Φ σ m 1 π m 2 .

Substitution in the above equation thus gives

R N G ( X , t ) d X = σ k 1 σ Ψ ( 2 π ) k 2 t k R k 0 1 cosh u m 2 G k 2 1 u Ψ 0 ( σ ) t u k 1 d u d σ = σ k 1 σ Ψ ( 2 π ) k 2 R k 0 1 cosh u m 2 G k 2 1 ( u Ψ 0 ( σ ) ) u k 1 d u d σ = ( 2 π ) k 2 R k 0 1 cosh u m 2 G k 2 1 ( u σ ) u k 1 d u d σ ,

where in the last equality, we have used Lemma 2.6 again. Compared with (4.3), we see that

(4.15) R N G ( X , t ) d X = ( 2 π ) k R k h ˆ σ 2 π d σ ,

where

h ( λ ) = 1 cosh λ m 2 .

Since h S ( R k ) , and h ( 0 ) = 1 , from the classical inversion formula for the Fourier transform, we find

R k h ˆ σ 2 π d σ = ( 2 π ) k R k h ˆ ( σ ) d σ = ( 2 π ) k h ( 0 ) = ( 2 π ) k .

Substituting in (4.15), we have proved (1.15), thus completing the proof of Theorem 1.1.□

5 The nonlinear heat equation sees the anisotropic Minkowski gauge Θ 0

This section is devoted to the following.

Proof of Theorem 1.3

To save on space, we will compute the integral

(5.1) I = 0 1 t m 2 + k 1 0 u sinh u m 2 e u tanh u Φ 0 ( z ) 2 4 t G k 2 1 u Ψ 0 ( σ ) t u k 1 d u d t t

and then multiply the result by the constant

(5.2) σ m 1 σ k 1 σ Φ σ Ψ ( 2 π ) k 2 ( 4 π ) m 2 .

To begin, we note that an obvious change of variable shows that

(5.3) I = 0 t m 2 + k 2 0 u sinh u m 2 e t u tanh u Φ 0 ( z ) 2 4 G k 2 1 ( u Ψ 0 ( σ ) t ) u k 1 d u d t = Ψ 0 ( σ ) 1 k 2 0 u sinh u m 2 u k 2 0 t m 2 + k 2 1 e t u tanh u Φ 0 ( z ) 2 4 J k 2 1 ( u Ψ 0 ( σ ) t ) d t d u ,

where the second inequality is justified by an exchange of the order of integration and by having used (1.13).

To unravel the inner integral in t on the right-hand side of (5.3), we now use the following formula due to Gegenbauer, see (3) on p. 385 in [22]: suppose that

(5.4) ( ν + μ ) > 0 , ( α + i β ) > 0 , ( α i β ) > 0 .

Then, one has

(5.5) 0 t μ 1 e α t J ν ( β t ) d t = 2 ν β ν Γ ( ν + μ ) Γ ( ν + 1 ) ( α 2 + β 2 ) ν + μ 2 F ν + μ 2 , 1 μ + ν 2 ; ν + 1 ; β 2 α 2 + β 2 ,

In (5.5), we have denoted by

(5.6) F ( α 1 , α 2 ; β 1 ; z ) = Γ ( β 1 ) Γ ( α 1 ) Γ ( α 2 ) k = 0 Γ ( k + α 1 ) Γ ( k + α 2 ) Γ ( k + β 1 ) k ! z k ,

Gauss’ hypergeometric function. We now assume z R m \ { 0 } , σ R k \ { 0 } , and for u ( 0 , ) fixed, apply (5.5) with the choice of parameters

ν = k 2 1 , μ = m 2 + k 2 , α = u tanh u Φ 0 ( z ) 2 4 , β = u Ψ 0 ( σ ) ,

which gives

ν + μ = m 2 + k 1 , 1 μ + ν = m 2 .

Note that, since m 1 , k 1 , we have ν + μ 1 2 > 0 , and since z 0 , all requirements in (5.4) are fulfilled. Since

α 2 + β 2 = u 2 16 tanh 2 u ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u ) , β 2 α 2 + β 2 = 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ ( σ ) 2 tanh 2 u ,

we obtain from (5.5)

(5.7) 0 t m 2 + k 2 1 e t u tanh u Φ 0 ( z ) 2 4 J k 2 1 ( u Ψ 0 ( σ ) t ) d t = 2 1 k 2 ( u Ψ 0 ( σ ) ) k 2 1 Γ m 2 + k 1 Γ k 2 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u ) 1 2 m 2 + k 1 × 16 tanh 2 u u 2 1 2 m 2 + k 1 F 1 2 m 2 + k 1 , m 4 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u = 4 m 2 + k 1 2 1 k 2 Ψ 0 ( σ ) k 2 1 Γ m 2 + k 1 Γ k 2 u k 2 1 ( tanh 2 u ) 1 2 m 2 + k 1 u m 2 + k 1 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u ) 1 2 m 2 + k 1 × F 1 2 m 2 + k 1 , m 4 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u .

Substituting (5.7) in (5.3), we find

(5.8) I = 4 m 2 + k 1 2 1 k 2 Γ m 2 + k 1 Γ k 2 0 1 sinh 2 u m 4 ( tanh 2 u ) 1 2 m 2 + k 1 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u ) 1 2 m 2 + k 1 × F 1 2 m 2 + k 1 , m 4 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u d u .

It might be helpful for the reader to note that in the above computation a first little miracle has happened: the powers of u have disappeared from the integrand. This crucial aspect hides an important geometric information. To proceed in the computation, we somehow need to kill the enemy, i.e., the very unpleasant factor

1 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u ) 1 2 m 2 + k 1

in the integral (5.8). The appropriate tool for the job is the following Kummer’s relation (one of many), which prescribes the change of the hypergeometric function F under linear transformations (see formula (3) on p. 105 in [8]),

(5.9) F ( α , β ; γ ; x ) = ( 1 x ) α F α , γ β ; γ ; x x 1 , x 1 , arg ( 1 x ) < π .

Comparing (5.8) with (5.9), it appears evident that we should apply the latter to the choice

x x 1 = 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u , γ = k 2 , γ β = m 4 , α = 1 2 m 2 + k 1 .

This choice gives

x = 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 , β = k 2 + m 4 .

What is crucial for us is that

1 x = Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 0 ,

and therefore

( 1 x ) α = Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 1 2 m 2 + k 1 .

Applying (5.9), we thus find

(5.10) F 1 2 m 2 + k 1 , m 4 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u = Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 1 2 m 2 + k 1 × F 1 2 m 2 + k 1 , m 4 + k 2 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 .

Substitution of (5.10) into (5.8) gives

(5.11) I = 4 m 2 + k 1 2 1 k 2 Γ m 2 + k 1 Γ k 2 Φ 0 ( z ) 2 m 2 + k 1 0 1 sinh 2 u m 4 ( tanh 2 u ) 1 2 m 2 + k 1 × F 1 2 m 2 + k 1 , m 4 + k 2 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 d u .

Formula (5.11) represents the second little miracle in the proof of Theorem 1.3. Our next objective is to finally compute in closed form the integral on the right-hand side of (5.11). With this in mind, we intend to make the crucial change of variable y = tanh 2 u , which gives d y = 2 tanh u cosh 2 u d u . To do this, we rearrange part of the integrand in the following way:

1 sinh 2 u m 4 ( tanh 2 u ) 1 2 m 2 + k 1 = 1 2 1 cosh 2 u m 4 1 ( tanh 2 u ) k 2 2 2 tanh u cosh 2 u .

We need to be a bit careful here and distinguish the case k = 1 from k 2 , but the relevant details are left to the reader. Assuming that k 2 , we rewrite (5.11) as follows:

(5.12) I = 4 m 2 + k 1 2 1 k 2 Γ m 2 + k 1 2 Γ k 2 Φ 0 ( z ) 2 m 2 + k 1 0 ( 1 tanh 2 u ) m 4 1 ( tanh 2 u ) k 2 2 2 tanh u cosh 2 u × F 1 2 m 2 + k 1 , m 4 + k 2 ; k 2 ; 16 Ψ 0 ( σ ) 2 tanh 2 u Φ 0 ( z ) 4 d u .

Performing the above-stated change of variable in the integral on the right-hand side of (5.12), we finally reach the conclusion that

(5.13) I = 4 m 2 + k 1 2 1 k 2 Γ m 2 + k 1 2 Γ k 2 Φ 0 ( z ) 2 m 2 + k 1 0 1 ( 1 y ) m 4 1 y k 2 1 × F 1 2 m 2 + k 1 , m 4 + k 2 ; k 2 ; 16 Ψ 0 ( σ ) 2 Φ 0 ( z ) 4 y d y .

The form of the integral on the right-hand side of (5.13) represents the third miracle in the proof of the theorem. To unravel it, we apply the following formula due to H. Bateman, see [8, (2) on p. 78], which gives

(5.14) 0 1 ( 1 y ) γ c 1 y c 1 F ( α , β ; c ; a y ) d y = Γ ( c ) Γ ( γ c ) Γ ( γ ) F ( α , β ; γ ; a ) ,

provided that

γ > c > 0 , a 1 , arg ( 1 a ) < π .

Applying (5.14) with

γ = m 4 + k 2 , c = k 2 , α = 1 2 m 2 + k 1 , β = m 4 + k 2 , a = 16 Ψ 0 ( σ ) 2 Φ 0 ( z ) 4 ,

we finally obtain

(5.15) I = 4 m 2 + k 1 2 k 2 Γ m 2 + k 1 Γ m 4 Γ m 4 + k 2 Φ 0 ( z ) 2 m 2 + k 1 F 1 2 m 2 + k 1 , m 4 + k 2 ; m 4 + k 2 ; 16 Ψ 0 ( σ ) 2 Φ 0 ( z ) 4 .

If we now use the following elementary, yet important property, which can be directly verified from (5.6), see also [8, (4) on p. 101],

(5.16) F ( α , β ; β ; a ) = ( 1 + a ) α ,

we reach the remarkable conclusion that

(5.17) F 1 2 m 2 + k 1 , m 4 + k 2 ; m 4 + k 2 ; 16 Ψ 0 ( σ ) 2 Φ 0 ( z ) 4 = Φ 0 ( z ) 2 m 2 + k 1 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 ) 1 2 m 2 + k 1 .

Substituting (5.17) in (5.15), we have

(5.18) I = 4 m 2 + k 1 2 k 2 Γ m 4 Γ m 2 + k 1 Γ m 4 + k 2 ( Φ 0 ( z ) 4 + 16 Ψ 0 ( σ ) 2 ) 1 2 m 2 + k 1 .

To more conveniently represent the constant on the right-hand side of (5.18), we now use (3.14) in which we take x = 1 2 m 2 + k 1 , to find

Γ m 2 + k 1 Γ m 4 + k 2 = π 1 2 2 m 2 + k 2 Γ 1 2 m 2 + k 1 .

Substituting in (5.18), and keeping (1.24) and (1.25) in mind, we obtain

(5.19) I = π 1 2 2 3 2 m + 5 2 k 4 Γ m 4 Γ 1 2 m 2 + k 1 Θ 0 ( z , σ ) ( Q 2 ) .

If we now remember that we need to multiply the right-hand side of (5.19) by the constant in (5.2), we finally reach the conclusion that

(5.20) 0 G ( X , t ) d t = σ m 1 σ k 1 σ Φ σ Ψ 2 m 2 + 2 k 4 Γ m 4 Γ 1 2 m 2 + k 1 π m + k + 1 2 Θ 0 ( z , σ ) ( Q 2 ) .

In view of (3.10) in Proposition 3.5, equation (5.20) finally proves (1.26).□


Dedicated to Ermanno, with affection, on the occasion of his 80th birthday.


  1. Funding information: N. Garofalo is supported in part by a Progetto SID (Investimento Strategico di Dipartimento): “Aspects of nonlocal operators via fine properties of heat kernels,” University of Padova (2022); and by a PRIN (Progetto di Ricerca di Rilevante Interesse Nazionale) (2022): “Variational and analytical aspects of geometric PDEs.” He is also partially supported by a Visiting Professorship at the Arizona State University.

  2. Author contribution: The author confirms the sole responsibility for the conception of the study, presented results and manuscript preparation.

  3. Conflict of interest: The author states no conflict of interest.

References

[1] G. Akagi, K. Ishige, and R. Sato, The Cauchy problem for the Finsler heat equation, Adv. Calc. Var. 13 (2020), no. 3, 257–278. 10.1515/acv-2017-0048Search in Google Scholar

[2] F. Baudoin and N. Garofalo, Curvature-dimension inequalities and Ricci lower bounds for sub-Riemannian manifolds with transverse symmetries, J. Eur. Math. Soc. (JEMS) 19 (2017), no. 1, 151–219. 10.4171/jems/663Search in Google Scholar

[3] G. Bellettini and M. Paolini, Anisotropic motion by mean curvature in the context of Finsler geometry, Hokkaido Math. J. 25 (1996), no. 3, 537–566. 10.14492/hokmj/1351516749Search in Google Scholar

[4] L. Capogna, G. Citti, and G. Rea, A subelliptic analogue of Aronson-Serrin’s Harnack inequality, Math. Ann. 357 (2013), no. 3, 1175–1198. 10.1007/s00208-013-0937-ySearch in Google Scholar

[5] X. Cheng and Y. Feng, Harnack inequality and the relevant theorems on Finsler metric measure manifolds. ArXiv:2312.06404. Search in Google Scholar

[6] A. Cianchi and P. Salani, Overdetermined anisotropic elliptic problems, Math. Ann. 345 (2009), no. 4, 859–881. 10.1007/s00208-009-0386-9Search in Google Scholar

[7] F. Dragoni, N. Garofalo, G. Giovannardi, and P. Salani, A fundamental solution for a subelliptic operator in Finsler geometry. arXiv preprint arXiv:2401.06736. Search in Google Scholar

[8] A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi, Higher transcendental functions, Vol. I. Based on notes left by Harry Bateman. With a preface by Mina Rees. With a foreword by E. C. Watson. Reprint of the 1953 original, Robert E. Krieger Publishing Co., Inc., Melbourne, Fla., 1981, xiii+302 pp. Search in Google Scholar

[9] V. Ferone and B. Kawohl, Remarks on a Finsler-Laplacian, Proc. Amer. Math. Soc. 137 (2009), no. 1, 247–253. 10.1090/S0002-9939-08-09554-3Search in Google Scholar

[10] N. Garofalo, Unique continuation for a class of elliptic operators which degenerate on a manifold of arbitrary codimension, J. Differential Equations 104 (1993), no. 1, 117–146. 10.1006/jdeq.1993.1065Search in Google Scholar

[11] N. Garofalo, Fractional thoughts, New developments in the analysis of nonlocal operators, Contemporary Mathematics, 723, American Mathematical Society, Providence, RI, 2019, pp. 1–135.10.1090/conm/723/14569Search in Google Scholar

[12] N. Garofalo and G. Tralli, Heat kernels for a class of hybrid evolution equations, Potential Anal. 59 (2023), 823–856. 10.1007/s11118-022-10003-2Search in Google Scholar

[13] A. Grigor’yan, The heat equation on noncompact Riemannian manifolds, Matem. Sbornik 182 (1991), 55–87; Math. USSR-Sbornik 72 (1992), 47–77. Search in Google Scholar

[14] N. N. Lebedev, Special functions and their applications. Revised edition, translated from the Russian and edited by R. A. Silverman. Unabridged and corrected republication. Dover Publications, Inc., New York, 1972, xii+308 pp. Search in Google Scholar

[15] P. Li and S.-T. Yau, On the parabolic kernel of the Schrödinger operator, Acta Math. 156 (1986), no. 3–4, 153–201. 10.1007/BF02399203Search in Google Scholar

[16] B. Muckenhoupt and E. M. Stein, Classical expansions and their relation to conjugate harmonic functions, Trans. Amer. Math. Soc. 118 (1965), 17–92. 10.2307/1993944Search in Google Scholar

[17] S.-I. Ohta and K.-T. Sturm, Heat flow on Finsler manifolds, Comm. Pure Appl. Math. 62 (2009), no. 10, 1386–1433. 10.1002/cpa.20273Search in Google Scholar

[18] S.-I. Ohta and K.-T. Sturm, Bochner-Weitzenböck formula and Li-Yau estimates on Finsler manifolds, Adv. Math. 252 (2014), 429–448. 10.1016/j.aim.2013.10.018Search in Google Scholar

[19] L. Saloff-Coste, Uniformly elliptic operators on Riemannian manifolds, J. Diff. Geom. 36 (1992), 417–450. 10.4310/jdg/1214448748Search in Google Scholar

[20] R. Schneider, Convex bodies: the Brunn-Minkowski theory, Encyclopedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014, xxii+736 pp. Search in Google Scholar

[21] J. E. Taylor, Crystalline variational problems, Bull. Amer. Math. Soc. 84 (1978), no. 4, 568–588. 10.1090/S0002-9904-1978-14499-1Search in Google Scholar

[22] G. N. Watson, A treatise on the theory of Bessel functions. Reprint of the second (1944) edition. Cambridge Mathematical Library, Cambridge University Press, Cambridge, 1995, viii+804 pp. Search in Google Scholar

Received: 2024-02-26
Revised: 2024-08-07
Accepted: 2024-09-17
Published Online: 2024-11-11

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 5.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/agms-2024-0012/html
Scroll to top button