Abstract.
We prove Juhl type formulas for the curved Ovsienko–Redou operators and their linear analogues, which indicate the associated formal self-adjointness, thereby confirming two conjectures of Case, Lin, and Yuan. We also offer an extension of Juhl’s original formula for the GJMS operators.
1. Introduction
The GJMS operator of order is a conformally invariant differential operator with leading-order term defined on any Riemannian manifold of dimension , and this family generalizes the well-known second-order conformal Laplacian (also called Yamabe operator) and the fourth-order operator discovered by Paneitz [Paneitz1983]. The GJMS operators have been studied intensively during the past decades in connection with, for example, prescribed -curvature problems, higher-order Sobolev trace inequalities, scattering theory on conformally compact manifolds, and functional determinant quotient formulas for pairs of metrics in a conformal class.
Based on a theory of residue families, in a series of works [Juhl2009, Juhl2013], Juhl derived remarkable formulas that express GJMS operators as a sum of compositions of lower-order GJMS operators up to a certain second-order term or as linear combinations of compositions of second-order differential operators, through an ingenious inversion relation for compositions given credit to Krattenthaler [Juhl2013, Theorem 2.1]. Later, Fefferman and Graham [FeffermanGraham2013] provided an alternative proof of Juhl’s formulas, starting directly from the original construction on the ambient space but also requiring Krattenthaler’s insight. Juhl’s formulas have significant applications in the aforementioned study of GJMS operators, such as the asymptotic expansion of the heat kernel [Juhl2016] and prescribed higher-order -curvature problems [MazumdarVetois2020].
To study conformally covariant operators of rank three, Case, Lin, and Yuan [CaseLinYuan2022or] gave two generalizations of the GJMS operators;
here we focus on those that are formally self-adjoint.
The first is a family of conformally invariant bidifferential operators:
|
|
|
of total order .
They are called the curved Ovsienko–Redou operators because they generalize a family of bidifferential operators constructed by Ovsienko and Redou [OvsienkoRedou2003] on the sphere. Let
|
|
|
with the multinomial coefficient wherein . The operators are determined ambiently by
|
|
|
on ;
in this paper, tensor products are over . The second generalization is a family of conformally invariant differential operators:
|
|
|
of order associated with a scalar Weyl invariant of weight . These are determined ambiently by a scalar Riemannian invariant of weight as
|
|
|
|
on where
|
|
|
We refer the reader to Section 2 for an explanation of our notation and a description of how the ambient formulas determine conformally invariant operators.
In a recent work, Case and the second named author [CaseYan2024] proved the formal self-adjointness of the two families of operators, thereby answering in the affirmative two conjectures of Case, Lin, and Yuan [CaseLinYuan2022or]. Taking as an example, their main idea is to factorize on the Poincaré space and realize the Dirichlet form of as the coefficient of the logarithmic term in the Dirichlet form of . The Divergence Theorem then yields the desired formal self-adjointness. This method avoids the lack of either an equivalent description of these operators as an obstruction to solving some second-order PDE, on which the Graham–Zworski argument [GrahamZworski2003] highly relies, or the complicated combinatorial arguments required by Juhl in [Juhl2013].
However, the explicit structures of these two families of operators are not revealed through the arguments in [CaseYan2024], especially in view of Juhl’s formulas for the GJMS operators [Juhl2009, Juhl2013]. To understand the underlying formal self-adjointness for the operators and in a more direct way and to read the two operators from a more general setting, the main purpose of this paper is to prove the following two Juhl type formulas:
Theorem 1.1.
Let be a conformal manifold.
Let and if is even, we assume additionally that . Writing , the operator satisfies
(1.1) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative sequences , and such that
|
|
|
In particular, is formally self-adjoint.
Theorem 1.2.
Let be a conformal manifold and let be an ambient scalar Riemannian invariant. Let and if is even, we assume additionally that with the Kronecker delta. Then the operator satisfies
(1.2) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative integers and all sequences and of nonnegative integers such that
|
|
|
In particular, is formally self-adjoint.
In the above theorems, the additional condition for the even case is imposed to ensure that the operators and , respectively, are independent of the ambiguity of the ambient metric. This will be discussed in Section 2.
1.1. Outline of the idea
1.1.1. Fefferman and Graham’s argument
In [FeffermanGraham2013], Fefferman and Graham gave a direct proof of Juhl’s formula for the GJMS operators, starting from the original construction on the ambient space. In this subsection, we sketch their main idea and show how to improve their arguments to work with the differential operators and .
Given a Riemannian manifold , let be its straight and normal ambient space defined in (2.1). We set
|
|
|
and denote . Since the construction of GJMS operators is independent of the extension of on , Fefferman and Graham reformulated them via instead of . Through a direct computation [FeffermanGraham2013, eq. (2.4)], one has, for ,
(1.3) |
|
|
|
It is notable that is a second-order, formally self-adjoint operator on for each and we may regard it as the generating function
|
|
|
for a family of second-order, formally self-adjoint operators on . Now setting
|
|
|
we may formulate the GJMS operators as
(1.4) |
|
|
|
With recourse to a nice combinatorial argument, Fefferman and Graham proved the following result [FeffermanGraham2013, eq. (2.5)]:
Theorem 1.3.
For and additionally if is even,
(1.5) |
|
|
|
|
|
|
|
|
where the summation runs over all sequences of nonnegative integers such that
|
|
|
In particular, is formally self-adjoint.
Here the main problem is to compute the coefficient for each composition of -operators on the right-hand side of (1.5). To do so, we fix the to be worked with and start with a truncated composition with . To produce the desired -composition, we shall look at terms containing a -truncation with in the expansion of our selected truncated composition of -operators. These terms are multiples of the -truncation times a power of . Note that amongst the -operators in the truncated composition, it is clear that the zeroth term has contributed exactly times and the differentiation on has contributed times, thereby implying that the aforementioned power of has an exponent . Finally, to compute the coefficients associated with the -truncations in question, Fefferman and Graham cleverly showed that these coefficients satisfy a family of recursive relations [FeffermanGraham2013, eq. (3.5)] by a remarkable combinatorial argument, and hence arrived at (1.5) by solving these recursions.
When it comes to the curved Ovsienko–Redou operators , it can be shown that
(1.6) |
|
|
|
|
|
|
|
|
|
|
|
|
where we set . However, this time we cannot proceed with the argument of Fefferman and Graham since an analogous family of recursions becomes out of reach, mainly due to the twisted inner layer of compositions. Likewise, for the operator , we have, as formulated in [CaseLinYuan2022or, eq. (6.2)],
(1.7) |
|
|
|
|
|
|
|
|
The inserted ambient scalar Riemannian invariant also kills the expected recursive relations.
1.1.2. Casting the “Diffindo” charm
To overcome the issue caused by the lack of necessary recursions, we need to cast the charm of Diffindo for the two operators and . That is, we shall separate the analyses of the inner and outer layers of operator compositions.
We begin with the inner layer. Let . For and , we consider
|
|
|
It is notable that reduces to the GJMS operator by choosing and taking . By the definition of the -operators, we see that for each , the expansion of is a linear combination of a nonnegative power of times a composition of -operators acted on . The takeaway from our analysis is that by using an evaluation of a hypergeometric series (instead of looking for recursions), the coefficients in this linear combination can be explicitly expressed, as shown in Corollary 3.2.
Next, we continue with the outer layer. Note that the essential contributions of the inner layer are of the form for and for where . The analysis for the former can be copied from that for the inner layer, while the study of the latter, as given in Corollary 3.4, is much more complicated, relying on a trick for Lemma 3.3.
To finalize our arguments, we shall look not only at the operators and , but also at their generalizations with a few more free parameters added, as given in (1.16) and (1.19), respectively. The main advantage of these free parameters is that the application of induction becomes possible. By further utilizing the combinatorial identities shown in Appendix A, we finally arrive at the explicit expressions of the two families of generalized operators presented in Theorems 5.1 and 6.1. In particular, as pointed out in Corollaries 5.2 and 6.2, the nature of formal self-adjointness of the operators and is exclusive among the two generic families, making our Juhl type formulas more meaningful.
1.1.3. Notation and basic properties
To facilitate our analysis, we split the -operators by defining for :
|
|
|
|
|
|
|
|
where and .
Let . For the inner layer, we have introduced the operators for and :
(1.8) |
|
|
|
|
Meanwhile, we define
(1.9) |
|
|
|
If we expand in terms of the operators and , then all its terms are of the form
(1.10) |
|
|
|
where takes either or .
Hence, we may record the terms in the expansion of as
(1.11) |
|
|
|
where and are sequences of nonnegative integers with the length of one more than that of such that
(1.12) |
|
|
|
It is also notable that the omitted index of the -operators should be determined by its position in . Meanwhile, in the expansion of , we need to attach to the term a coefficient:
(1.13) |
|
|
|
|
|
|
|
|
To study the curved Ovsienko–Redou operators , we look at a generic family of operators:
(1.14) |
|
|
|
where . Furthermore, we write
(1.15) |
|
|
|
It is clear that the curved Ovsienko–Redou operator is a specialization of
(1.16) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where and are fixed nonnegative integers and , and are indeterminates.
For the operators , we look at another generic family of operators:
(1.17) |
|
|
|
|
where and . Also, we write
(1.18) |
|
|
|
Now can be generalized as
(1.19) |
|
|
|
|
|
|
|
|
|
|
|
|
where and are fixed nonnegative integers and and are indeterminates.
3. Diffindo
In Subsection 1.1.2, we have pointed out that the key in our argument is to split the analyses of the inner and outer layers of operator compositions in and . Particularly, what are produced from the inner layer are essentially or with and . Since the outer layer is simply of the form , we shall look at and to cast the Diffindo. In view of (1.11), we first need to evaluate and for arbitrary sequences and of nonnegative integers such that
|
|
|
3.1. Evaluation of
We begin with . To simplify our notation, we write
(3.1) |
|
|
|
The following result gives an explicit expression of .
Lemma 3.1.
For any nonnegative integer ,
(3.2) |
|
|
|
|
(3.7) |
|
|
|
|
In addition, vanishes if there exists a certain index such that
(3.8) |
|
|
|
Proof.
For arbitrary and , we note that
|
|
|
Hence,
|
|
|
|
|
|
|
|
Repeatedly applying the above argument yields (3.2).
For the second part, it is trivial when since in this case no operator is acted on . For , let be the smallest index such that (3.8) holds. It follows that
|
|
|
Furthermore, if ,
|
|
|
Otherwise, the fact that is the smallest index ensuring (3.8) implies
|
|
|
Hence, we have the vanishing of the binomial coefficient
|
|
|
thereby implying that also vanishes.
∎
The above evaluation immediately leads us to an explicit expression of for independent of , which further gives by taking . More importantly, this result, as shown in the next corollary, serves as a generalization of Juhl’s formula in Theorem 1.3.
Corollary 3.2.
The operator satisfies
(3.9) |
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all sequences of nonnegative integers such that
(3.10) |
|
|
|
Proof.
We start by writing as
|
|
|
|
|
|
|
|
In view of (1.12), is such that
(3.11) |
|
|
|
In addition, since , we further require that the power of in reduces to zero. By (3.2), we have
|
|
|
so that satisfies
(3.12) |
|
|
|
Running over nonnegative sequences and restricted as above and invoking (3.2), we have
|
|
|
|
|
|
|
|
|
For the inner summation on , we make use of (A.6) with the substitutions:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Hence,
|
|
|
|
|
|
where we have also used (3.11) and (3.12). This finishes the proof of (3.9).
∎
3.2. Evaluation of
Now we consider for general .
Lemma 3.3.
For any nonnegative integer and any smooth function ,
(3.13) |
|
|
|
|
|
|
|
|
(3.18) |
|
|
|
|
where is the -th order derivative with respect to . In addition, the above summand vanishes for all .
Proof.
For arbitrary and ,
|
|
|
We observe that after applying the -operator, the derivative order of changes by , and accordingly the power of changes by . In , we are to have applications of , and hence the derivative order of ranges over the interval . This, in particular, confirms the second part of our result.
Now we write
(3.19) |
|
|
|
For the evaluation of the coefficients , our trick is to choose . The takeaway is that the expression
|
|
|
equals when , and vanishes for all other . Hence,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Invoking (3.2) gives the desired relation.
∎
As a consequence, we also arrive at an explicit expression of for independent of , which reduces to by taking .
Corollary 3.4.
For any nonnegative integer and any smooth function ,
(3.20) |
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative integers and all sequences of nonnegative integers such that
(3.21) |
|
|
|
Here by abuse of notation, we read as .
Proof.
Note that can be written as
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we have utilized (3.19). Still, is such that
(3.22) |
|
|
|
Meanwhile, we have shown in the proof of Lemma 3.3 that
|
|
|
Thus, if is such that
|
|
|
we must have the vanishing of for all by the second part of Lemma 3.1, and thus the vanishing of . Now for , it suffices to restrict
(3.23) |
|
|
|
so that and satisfy
(3.24) |
|
|
|
With the additional restriction for in (3.23), we further find that when ,
|
|
|
which, in light of the second part of Lemma 3.1, implies the vanishing of for these . Hence, for restricted by (3.23), we always have
|
|
|
|
|
|
|
|
by invoking (3.2). Now running over nonnegative integers and nonnegative sequences and as restricted by (3.22) and (3.24), we have
|
|
|
|
|
|
|
|
|
Note that the inner summation on is exactly the one in the proof of Corollary 3.2 with replaced by . The claimed result therefore follows.
∎
5. Formal self-adjointness of
Recall from Section 2 that the curved Ovsienko–Redou operators are determined ambiently by
|
|
|
where, in light of Lemma 2.4,
|
|
|
with
|
|
|
Here
|
|
|
For convenience, we further set
|
|
|
which simplifies as
|
|
|
In view of (1.3), we rewrite the -operators in terms of the -operators and arrive at
(5.1) |
|
|
|
|
|
|
|
|
|
This was already presented in (1.6). It is also clear that can be specialized from
|
|
|
|
|
|
|
|
|
|
|
|
by observing that
(5.2) |
|
|
|
To expand , we begin with the operators
|
|
|
Recall that and are smooth functions independent of and let and be nonnegative integers.
For the inner layer, we have two multiplicands, and we evaluate them separately. By (1.13) and (3.2),
|
|
|
|
|
|
|
|
|
|
|
|
in which the nonnegative sequence is such that
(5.3) |
|
|
|
in light of (1.12). Similarly,
|
|
|
|
|
|
|
|
|
|
|
|
where the nonnegative sequence is such that
(5.4) |
|
|
|
For the outer layer, we are essentially looking at
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
in view of (3.9). Here the nonnegative sequences , and are such that
|
|
|
according to (3.10). This is, in light of (5.3) and (5.4), further equivalent to
(5.5) |
|
|
|
For convenience, we write
(5.6) |
|
|
|
|
|
|
|
|
It follows that
(5.7) |
|
|
|
|
|
|
|
|
|
|
|
|
(5.12) |
|
|
|
(5.17) |
|
|
|
where , , , and are controlled by (5.3), (5.4) and (5.5).
Now we are in a position to expand .
Theorem 5.1.
Let and be nonnegative integers. Then
(5.18) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative sequences , and such that
(5.19) |
|
|
|
Proof.
We apply our previous analysis to expand each
|
|
|
First, we note that the restriction (5.19) comes from (5.5) where we have also utilized the fact that . Next, it is easy to observe that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Also,
|
|
|
Hence, equals
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Now we extend the sequence to the following sequence of length :
|
|
|
It is clear that
|
|
|
|
|
|
|
|
where we have applied (5.4) for the second equality. In (A.8), we replace with and make the following substitutions:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
It follows that
|
|
|
|
|
|
equals
|
|
|
|
|
|
Therefore, further equals
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
This time we extend the sequence to the following sequence of length :
|
|
|
Then by (5.3),
|
|
|
|
|
|
|
|
Let us replace with in (A.8) and make the following substitutions:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Thus,
|
|
|
|
|
|
equals
|
|
|
|
|
|
Substituting this relation into the expression of derived earlier yields the desired result.
∎
We are interested in for which choice of parameters the operator
|
|
|
is formally self-adjoint. Let be the coefficient of the monomial
|
|
|
|
|
|
Then these coefficients should satisfy
|
|
|
|
|
|
|
|
Recalling also that , and are nonnegative integers, the above imply that , and .
Corollary 5.2.
We have
(5.20) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative sequences , and such that
|
|
|
In particular, is formally self-adjoint.
Finally, we recall that
|
|
|
and hence complete the proof of Theorem 1.1.
6. Formal self-adjointness of
According to (1.7), the operator is a specialization of
|
|
|
|
|
|
|
|
by taking , , and replacing with .
In view of this, the main objective of this section is to derive an explicit expansion of the operator . To achieve this goal, we first need to look into the operators
|
|
|
Recall that is a smooth function independent of and that is a nonnegative integer.
For the inner layer, we may use (1.13) and (3.2) to get
|
|
|
|
|
|
|
|
|
|
|
|
Here the nonnegative sequence is such that
(6.1) |
|
|
|
which comes from (1.12).
For the outer layer, we are essentially looking at
|
|
|
|
|
|
|
|
|
|
|
|
in view of (3.20). Here the nonnegative integer and nonnegative sequences and are such that
|
|
|
according to (3.21). This is, in light of (6.1), further equivalent to
(6.2) |
|
|
|
It follows from the above discussion that
(6.3) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(6.8) |
|
|
|
|
where , , and are controlled by (6.1) and (6.2).
Now we are ready to produce an explicit expression of .
Theorem 6.1.
Let be a nonnegative integer. Then
(6.9) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative integers and all sequences and of nonnegative integers such that
(6.10) |
|
|
|
Here by abuse of notation, we read as .
Proof.
We directly use the previous analysis to evaluate each
|
|
|
Firstly, the restriction (6.10) simply comes from (6.2). Meanwhile, by (6.3), we know that
|
|
|
equals
|
|
|
|
|
|
|
|
|
where
|
|
|
|
|
|
|
|
In view of the outer summation on , we are left to analyze
|
|
|
Recall from (6.1) that is such that
|
|
|
Hence, we extend to a new sequence of length :
|
|
|
In particular,
|
|
|
|
|
|
|
|
Writing the summation in terms of , we see that
|
|
|
|
|
|
equals
|
|
|
|
|
|
where we have applied (A.8) with the substitutions:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Thus, our proof is complete.
∎
Now we consider the case that is formally self-adjoint. In this circumstance, the coefficient of each monomial
|
|
|
should satisfy , which implies that , and .
Corollary 6.2.
We have
(6.11) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the summation runs over all nonnegative integers and all sequences and of nonnegative integers such that
|
|
|
In particular, is formally self-adjoint.
The above result matches the shape of the operators according to our earlier discussion:
|
|
|
thereby closing the proof of Theorem 1.2.