Exponential decay for damped Klein-Gordon equations on asymptotically cylindrical and conic manifolds
Abstract.
We study the decay of the global energy for the damped Klein-Gordon equation on non-compact manifolds with finitely many cylindrical and subconic ends up to bounded perturbation. We prove that under the Geometric Control Condition, the decay is exponential, and that under the weaker Network Control Condition, the decay is logarithmic, by developing the global Carleman estimate with multiple weights.
1. Introduction
In this paper we study the decay of the global energy for the damped Klein-Gordon equation (1.6), on non-compact manifolds with finitely many ends of a wide class up to bounded perturbation, described in (1.1), including asymptotically cylindrical and conic ends. We prove in Theorem 1 that under the Geometric Control Condition given by Definition 1.2, in which the average of damping along each geodesic is uniformly bounded from below, the global energy decays exponentially. We prove in Theorem 2, that under the Network Control Condition given by Definition 1.3, in which each point in the space is within some uniform distance from the sufficient damped region, the global energy decays logarithmically. These results generalise those in [bj16]. The main new tool is the Carleman estimates with multiple weights in Theorem 3.
1.1. Geometric setting
Consider the model manifold , a non-compact connected -dimensional manifold without boundary, with infinite ends,
(1.1) |
where is a compact, connected manifold with boundary . Denote the interior of by . Each end is identified as a cylinder endowed with a product metric
(1.2) |
where is a -dimensional compact manifold without boundary, a smooth metric on . The scaling functions satisfy either one of the following conditions:
(1.3) | |||
(1.4) |
We call ends with the scaling functions in (1.3) sub-conic ends, and ends with those in (1.4) cylindrical ends. Specifically, sub-conic ends with the scaling function are called conic ends.
In this paper, we specify and work with bounded perturbations of the model metric (1.2). Specify a manifold of bounded geometry that is a bounded perturbation of our model manifold , in the sense that both the identity map
(1.5) |
and its inverse are -maps between two manifolds. See Appendix LABEL:SA for the definition of boundedness on manifolds of bounded geometry. Note that is uniformly bounded at each , from both above and below as a map from equipped with to equipped with .
We inexhaustively list some examples that are compatible with our setting:
Example 1.1.
-
(1)
Euclidean spaces , with being the unit open ball and being the rest of as in spherical coordinates. This is a conic end with .
-
(2)
Euclidean spaces as above, but endowed with a bounded perturbed metric, whose local matrix form in the canonical Euclidean coordinates, , and its inverse , are smooth matrix-valued functions of which components are .
-
(3)
Asymptotically conic manifolds, also as known as Riemannian scattering spaces, of finitely many ends of the form endowed with scattering metrics . Here ’s are smooth symmetric 2-cotensors on whose restriction to is positive-definite. In our model we realise the metric as a bounded perturbation of where and ’s are metrics on independent of . See [mel95] for further details.
-
(4)
Product cylinders of the form where is a closed manifold, by taking , and . Or an one-ended cylinder glued to some closed manifold. More generally, asymptotically cylindrical manifolds also work with our setting. Those are manifolds with finitely many ends of the form endowed with . Here again is a smooth symmetric 2-cotensor on whose restriction to is positive-definite. See [mel95] for further details.
-
(5)
Elliptic paraboloid, with being the tip and be the rest of paraboloid as equipped with metric , under the change of coordinates . Here this metric on is a bounded perturbation of , whose scaling function is , which is sub-conic.
-
(6)
Boundedly perturbed cylinders. Consider the surface , which is realised as the bounded perturbation to equipped with metric . Note that, this surface is not an asymptotically cylindrical manifold, as the cannot be well-defined at the spatial infinity . But we could still cope with this manifold as a bounded perturbation of the product cylinder.
-
(7)
Any connected sum along balls of finitely many equidimensional ends of the types above.
In this paper, we prove that in the geometric settings as above, one has exponential or logarithmic decays of the global energy for the damped Klein-Gordon equations by assuming suitable dynamical control conditions. It is also noted that, hyperbolic manifolds do not fit our analysis.
1.2. Damped Klein-Gordon equations
Consider a damping function , a smooth function on whose derivatives of all orders are bounded by uniform constants dependent only on the order. The damped Klein-Gordon equation on our manifold reads
(1.6) |
where is the positive Laplace-Beltrami operator on . Consider
(1.7) |
which is a bounded linear operator from to , which is further dissipative in the sense that
(1.8) |
By noting that is dense in , is a bounded dissipative operator on Hilbert space , and the Lumer-Phillips theorem tells us generates a strongly continuous semigroup on that is further a contraction semigroup, in the sense that for each . Note we can formulate the equation (1.6) as a Cauchy problem for , that is
(1.9) |
and the strongly continuous semigroup is the solution operator to the Cauchy problem, where the unique solution is . As we look into how fast the global energy decays, it suffices to look at the decay of the operator norm of the semigroup . Indeed, the energy of the solution to (1.6) is
(1.10) |
The semigroup weakly decays to 0 when the damping is smooth and not zero somewhere on , as in [wal77]. In this paper, we are interested in two types of decays of the semigroup : the exponential decay, which is the fastest decay one expects in the contexts of a smooth and bounded damping, and the logarithmic decay, which is a kind of non-uniform decay under some weak dynamical hypotheses.
About the damped Klein-Gordon equation and the damped wave equation, there have been many results known when is compact and the damping is smooth. It is known that the exponential decay of the semigroup is equivalent to the Geometric Control Condition, which is a dynamical hypothesis that all trajectories of the Hamiltonian flow intersect the support of the damping , as in [rt74, blr88, blr92, bg97]. In [leb93] it was shown that there is a logarithmic decay as long as the damping is non-trivial. It is also noted that other non-uniform stability properties have been actively investigated, as in, inexhaustively listed here, [aln14, bh07, csvw14, bc15].
However the picture is less complete for exponential results on non-compact manifolds without boundary. The fundamental result in [bj16] generalises the Geometric Control Condition to , with a uniform lower bound of the average of damping along the Hamiltonian flow. It was also shown that the Geometric Control Condition gives exponential decay of the semigroup, and that there is logarithmic decay when another dynamical hypothesis, called the Network Control Condition, is imposed. In [wun17] a polynomial decay was shown via Schrödinger observability for a periodic damping on under no further dynamical assumptions. In [jr18], the sharp polynomial global energy decay for the damped wave equation on with an asymptotically periodic damping was shown. In [roy18a] the results of [bj16, wun17] were extended to highly oscillatory periodic dampings. See also, inexhaustively listed here, [mr18, ms18, gjm19, gre19, cpsst19] for recent development on Euclidean spaces.
The purpose of this paper is to extend the results of [bj16] to a wider class of open manifolds, namely prespecified in (1.5). In [bj16], the results have been shown for the Euclidean cases (1), (2) of Example 1.1. The possibility of proving such results on product cylinders as in (4) was also hinted. Our paper generalises their results to manifolds with cylindrical and sub-conic ends. Here we define the Geometric Control Condition on the prespecified manifold :
Definition 1.2 (Geometric Control Condition).
We say the damping satisfies the Geometric Control Condition on , for , if for , where , one has
(1.11) |
where is the Hamiltonian flow associated with , and is the projection from fibres of to the base variable.
We claim the first main result that Geometric Control Condition gives exponential decay of the semigroup :
Theorem 1 (Exponential decay of energy).
Assume where everywhere, satisfies the Geometric Control Condition , then the semigroup decays exponentially in the sense that , for each , for some . It is then implied that the solution to the damped Klein-Gordon equation with initial datum ,
(1.12) |
decays exponentially, in the sense that there exists ,
(1.13) |
The decay of the global energy for the damped Klein-Gordon equation is determined by how fast the high frequency waves and the low frequency waves decay. The energy of the high frequency waves semiclassically concentrates near the Hamiltonian flow. This phenomenon hints at why the Geometric Control Condition plays an important role here. On the other hand, the low frequency waves do not concentrate. But as a result of their long wavelengths, they can see the damping from a distance even if many trajectories do not encounter the damping. However, the sparser the damping is, the weaker the decay gets. Therefore to obtain a uniform rate of decay we do not want to be too far away from the damping. This inspires the following dynamical hypothesis.
Definition 1.3 (Network Control Condition).
We say the damping satisfies the Network Control Condition on , for , and a set of points on , if at each ,
(1.14) |
and on .
This hypothesis has been introduced in [bj16], and gives logarithmic decay on . The logarithmic decay on compact manifolds has also been considered in [leb93, lr97]. Here is our second result:
Theorem 2 (Logarithmic decay of energy).
Assume where everywhere, satisfies the Network Control Condition , then for each , the solution to the damped Klein-Gordon equation with initial datum ,
(1.15) |
decays logarithmically, in the sense that there exists ,
(1.16) |
Though the idea of the proof is similar to that of [bj16], we need new tools because of we are leaving . In [bj16], they used the fact that has critical points exactly at . On our , neither the function nor the -structure remains. We manage to get this fixed on cylindrical ends, but it remains unfixable on those sub-conic ends.
To counter such difficulty in dealing with the subconic ends, we develop a novel Carleman estimate using a finite family of weight functions on manifolds of bounded geometry without boundary. The idea is based on that of two-weight Carleman estimates in bounded domain developed in [bur98]. The new estimate allows us to construct on each end a finite family of Carleman weights, possibly very degenerate or even identically a constant somewhere, to cover the whole manifold and to give a global Carleman estimate. To our knowledge, this global Carleman estimate using finitely many weight functions has not been employed previously. This Carleman estimate with multiple weights might be interesting on its own for other applications.
We note here that the regularity assumptions upon the damping in these two theorems can be weakened. In Theorem 1 we only need to be uniformly continuous, and in Theorem 2 we only need . We choose not to develop those improvements here but they follow from the strategy described in [bj16].
We organise our paper in the following order: in Section 2, we introduce our Carleman estimate with multiple weights; in Section 3, we show there exists a family of Carleman weight functions on our prespecified manifold compatible with the Carleman estimate developed in Section 2; in Section 4, we finish the proof of Theorem 1 concluding the exponential decay; in Section 5, we finish the proof of Theorem 2 concluding the logarithmic decay. An appendix on analysis on manifolds of bounded geometry is attached at the end of the paper.
1.3. Acknowledgement
The author is grateful to Jared Wunsch for numerous discussions around these results as well as many valuable comments on the manuscript, and Nicolas Burq for helpful discussion and pointing out the possibility to use the two-weight Carleman estimate, and Jeffrey Rauch and Jacob Shapiro for their insightful comments. The author is grateful to two anonymous referees for kindly reading this manuscript and providing many valuable remarks.
2. Carleman estimates with multiple weights
Let be a manifold of bounded geometry, without boundary. See Appendix LABEL:SA for further details. Let be an open set.
Definition 2.1 (Compatibility conditions).
We say a finite family of weight functions is compatible with control from , if there exists an open set with the following properties:
-
(1)
We have where is the distance on .
-
(2)
There exist constants such that, at each point , for each , if , then there exists some that
(2.1)
It is natural to impose the compatibility condition upon the weight functions. We aim to control the -size of a quasi-mode by merely the -size of that inside the region of control . At outside the region of control, if we allow some weight functions to have vanishing gradients, they will not control the size of the quasi-mode locally near . Therefore there has to be another weight whose gradient is sufficiently large to control that locally near . This explains the first part of (2.1).
On another hand, at such a point , because we use the exponential weights whose control is exponentially weak, we do not want this very weak control to be cloaked by the large exponential sizes of other non-controlling weights . To avoid that, we ask for a fixed gap between non-controlling and controlling weights, as in the second part of (2.1). Then we have for some uniform constant depending on . Now we note that the control induced by is observable, in the sense that the non-controlling weight generates an exponentially weaker term.
Theorem 3 (Global Carleman estimates with multiple weights).
For a manifold of bounded geometry without boundary, assume there are non-negative Carleman weights compatible with the control from in the sense of (2.1). Then, we have a global Carleman estimate with constant , independent of semiclassical parameter for small , such that
(2.2) |
where is a semiclassical uniformly bounded real potential.
Proof.
1. We start by deriving a local estimate via the hypoelliptic arguments. First note we can write
(2.3) |
where and . Fix a cutoff such that on and identically on . Fix a and denote . For each , fix a such that on and identically on . Set . Construct the exponential Carleman weights by , where is some large number to be determined later, and the conjugated operator by
(2.4) |
See (LABEL:A1L8) for the notation . Note
(2.5) |
where and the real and imaginary parts of the principal symbol are
(2.6) |
Denote the subset of the cotangent bundle that contains the characteristic set,
(2.7) |
outside of which . Consider a microlocal cutoff that is supported in and is identically 1 on . Note and have the same principal symbol, so
(2.8) |
On another hand, let then and are both in and their principal symbols agree. Thus
(2.9) |
From (2.8) and (2.9) we know that
(2.10) |
We claim that is uniformly bounded from below on . Note that at any we have from the definition of . Hence on we have and therefore
(2.11) |
where ’s are some constants dependent only on the maximal size of first derivatives of ’s. Let . In the canonical coordinates induced by the geodesic normal coordinates around , we have and . Compute, by noting that at we have , the Euclidean gradient,
(2.12) |
and
(2.13) |
from the uniform boundedness of the derivatives of order up to 2 of . Hence on with a large ,
(2.14) |
We conclude that throughout , we have
(2.15) |
for some fixed large, and hence there exists such that
(2.16) |
Now by invoking the weak Garding inequality in Proposition LABEL:A1T1 on we have for any with support inside ,
(2.17) |
This implies
(2.18) |
From (2.5), we use (2.8) and (2.10) to obtain
(2.19) |
and absorb the last term: for any with support inside we have
(2.20) |
Apply the above estimate to to obtain the local estimate,
(2.21) |
from the claimed hypoellipticity.
2. We want to derive a crude version of the global estimate by just summing up the local estimates. Let (2.21) be further simplified. Estimate
(2.22) |
and
(2.23) |
Let be a cutoff function supported inside and being identically on the support of , and let be cutoff functions supported inside and identically on the support of . We immediately have the estimates of the first two terms in (2.23),
(2.24) |
from the uniform boundedness of first two derivatives of and . Consider the next term,
(2.25) |
The first term of the last line is estimated via an adjoint argument
(2.26) |
Indeed, we have
(2.27) |
We bring (2.25) and (2.26) together to see
(2.28) |
since the third order derivatives of are uniformly bounded and is a bounded potential. Symmetrically we have
(2.29) |
and a complete estimate of (2.23),
(2.30) |
Finally we have for ,
(2.31) |
Note that the constants could be chosen uniformly such that (2.31) holds for each . Sum up (2.31) over to get the crude version of the global estimate,
(2.32) |
in which is supported inside , and is supported inside .
3. Finally we use the compatibility condition (2.1) imposed upon the Carleman weights to refine the global estimate (2.32). Recall that stands for the points at which the weight fails to control the quasimode. Given the assumptions on compatibility (2.1), at each , there exists some such that
(2.33) |
where
(2.34) |
Note that does not depend on . We have
(2.35) |
Now at each , we can partition into and with some , where
(2.36) |
For each , as , there is some such that
(2.37) |
(2.38) |
As , the cutoffs ’s are all , and therefore
(2.39) |
for . Note that when , (2.39) holds trivially because all ’s are 1, and therefore we conclude that (2.39) holds at each . This improves the estimate (2.32) from the left:
(2.40) |
as is supported inside .
Meanwhile, for each , the cutoff function is supported in . Therefore at each , we know from (2.33) and (2.35) that
(2.41) |
This inequality outside holds trivially as vanishes, and hence holds everywhere in . It improves the second term on the right in (2.32),
(2.42) |
Bring (2.40) and (2.42) into (2.32) to observe
(2.43) |
Finally, as and are both supported inside and are uniformly bounded, we can bound the -norm of terms and by the -norm of , that is,
(2.44) |
As semiclassically, we can absorb the second term on the right by the term on the left for small , that is,
(2.45) |
Denote the global maximum and minimum over all ’s by
(2.46) |
where . Then we have from (2.45),
(2.47) |
which is reduced to
(2.48) |
This is our claim. ∎
Remark 2.2.
-
(1)
When is a compact manifold without boundary, then one could control from any open set with arbitrary open subset with only one weight. It suffices to find a Morse function and find a diffeomorphism moving all critical points into .
-
(2)
It is observed that the uniform gap in the compatibility condition is necessary. By this fixed gap , we extracted an -decay in (2.41), further leading to the absorption argument between (2.44) and (2.45). Without such this uniform gap we see the inequality (2.44) will not generate any effective bound on -norm of , as this term on the right is now of size semiclassically.
3. Construction of Carleman weights
In this section, we aim to explicitly construct the weight functions on our prespecified manifold in (1.5) to obtain the global Carleman estimate we developed in the previous section. Assume throughout this section that the Network Control Condition defined in Definition 1.3 holds on . Let . Our ultimate target in this section is to control the whole manifold from .
The strategy is to start by working on the model manifold in (1.1). We construct a family of weights on each end, and another weight on the central compactum, then show they are compatible with control from . Eventually we pull back the weights via back to the prespecified .
Note that for each , and . We claim that a Network Control Condition on implies another on . Recall that is bounded from above and below,
(3.1) |
and therefore
(3.2) |
for each . Let and . At each , we have
(3.3) |
the last inequality of which comes from the Network Control Condition on . We also have on as an immediate result of (3.2). Therefore we could, without loss of generality, assume a Network Control Condition on . Note that
(3.4) |
We begin by constructing weight functions on the cylindrical ends, where the scaling functions ’s are identically 1.
Lemma 3.1 (Cylindrical ends).
Consider a cylindrical end , that is endowed with the metric , where is a smooth metric on closed . There exists , where and there is some such that,
(3.5) |
Proof.
In this lemma we Let for we here only care what is happening on .
1. We start by constructing a prototype weight based on the periodic structure. On pick a Morse function that is positive on . As is compact, has critical points at . Fix small. Let a periodic function be given by
(3.6) |
for small . This is a function with a period of . Consider
(3.7) |
and modify its size to get
(3.8) |
where we note that and we will later modify to move around the critical points. The critical points of are
(3.9) |
where .
2. We modify the weight to have critical points of distance uniformly bounded from below by from each other. For , define the flows for by
(3.10) |
generated by the constant radial vector fields . Also denote by the flow ,
(3.11) |
that preserves the periodicity of , in the sense that for each . Note that the flow pulls back points
(3.12) |
to critical points , that is, . Let and to be the -neighbourhood of , for . Now for , construct diffeomorphisms with inverses such that
(3.13) |
Because that all ’s are disjoint and on , we can glue up ’s to obtain a diffeomorphism on . Note that have all derivatives uniformly bounded from above and below, as the domain is compact. Also note that is the identity on and . This enables us to extend periodically to some on , by defining on , for each ,
(3.14) |
Let , whose critical points are
(3.15) |
for each , each . Renumber those critical points by . We remark that any two critical points of are separated by distance of at least . We also note for each one has outside for constant only depending on , because is still periodic.
3. Finally we modify the weight function in uniform radius balls around critical points to obtain (3.5). Note that the balls
(3.16) |
for any pair of critical points and . By the Network Control Condition, in each ball we can find some in the network such that
(3.17) |
and on . Now in each ball , find a diffeomorphism such that
(3.18) |
Glue up ’s to get a diffeomorphism on . We remark here that we can make this construction uniform in the sense that both are in , as in [bj16, rm16]. Therefore we have for some uniform in all . Now set . We know uniformly for all . Hence for any , we have , and again by the boundedness of we have for some uniform . As in (3.4), , the claim holds.
∎
Remark 3.2.
What makes this construction above interesting is that it is global on each cylindrical end, similar to the flavour of that on in [bj16]. So it only takes a single weight function to control the whole end. However, it still relies much on the homogeneity of the space along the radial direction. Once we allow the scaling functions ’s to grow as , for example, on conic ends, this construction stops working, technically because there is no ideal way of constructing a product-type in (3.7). This constraint on subconic ends is removed by introducing a finite collection of weights.
Lemma 3.3 (Subconic ends).
Consider a sub-conic end , that is endowed with the metric , where is a smooth metric on closed , and as described in (1.3). There exists some and let , and for some finite there exist that each , with a constant , such that for all , at each point with , there is some depending on , such that and
(3.19) |
Proof.
In this lemma we write and for we here only care what is happening on .
1. We start by quasi-isometrically reducing the underlying geometry to an unbounded subset of . As is compact, it possesses a finite cover
(3.20) |
such that each is a -diffeomorphism. For convenience, denote for ,
(3.21) |
where consists of points of distance more than away from the unit sphere, is the -neighbourhood of the unit sphere, and is the unit open ball. Here, as the covers are open, one can fix a small such that for each , for each such that , there is some such that . Denote the model space by where
(3.22) |
Construct diffeomorphisms and observe that covers . Each map is uniformly quasi-isometric with constants such that
(3.23) |
for any . Consider for
(3.24) |
This is a quasi-isometric -diffeomorphism, that for any we have
(3.25) |
To verify the nature of , it suffices to first pull back to , and then verify that the Christoffel symbols on are on as a subset of . We omit the trivial computation here. Note that each is a -diffeomorphism from to quasi-isometric with constants .
2. Now we construct on each , a weight function with vanishing gradients exactly inside the damping balls given in the Network Control Condition. On we construct for ,
(3.26) |
whose critical points are
(3.27) |
for all . See Figure 3. Note that any two such critical points are of distance at least , measured in . Set be the smallest constant such that for all , we have . This lower bound on the radius guarantees that any point in is of distance larger than from the cross-sectional boundary , measured in . By the Network Control Condition, for all critical points ’s that are inside there exists at least a such that and on . Here
(3.28) |
are disjoint balls around ’s of some uniform radius. Hence via a process similar to the construction of the diffeomorphism in the proof of 3.1, we can find a -diffeomorphism on , equal to the identity on , such that . Set , whose critical points in are a subset of . Set . This is a function defined on . Note that .
3. We now very carefully cut off the part of within a small neighbourhood of , and pull back and extend it to weight functions on for some . Observe that away from one has for some small . Set to be that for all , . Construct a cross-sectional cutoff such that on , greater than on , less than on , and identically on . Moreover we ask . See Figure 4. Note that we can find such a cutoff because is taken large enough to give the cross-section enough space to accommodate the tempered decay. Let the weight functions on be . As is identically zero near , we extend to all of by . Note that in general , and specifically on we have .
4. We claim that ’s meet our requirement listed in the statement. There is a lower bound for the pushforward map . Hence we have
(3.29) |
Fix and some point while . Note this means and . Now set . There are three circumstances depending where is.
(a) If , the cross-sectional cutoff is identically . We have
(3.30) |
and therefore .
(b) If , the cross-sectional cutoff is sufficiently large. We have
(3.31) |
Here we used the fact that is bounded from above by 3. Therefore .
(c) If , there is some such that . From the circumstance (a), we know that , and
(3.32) |
Here we used the fact that and .
The claim has been concluded as above. ∎
Remark 3.4.
-
(1)
We note that the argument is sharp for conic ends, where . If is not uniformly bounded, then loses the quasi-isometric nature, and the argument needs further modifications.
-
(2)
This argument relies, twice when setting up and , on the fact that the cross-sectional space is expanding as . Large makes sure that the is sufficiently apart from , so the critical points inside will not be pulled to some out of the charted region . As in the cylindrical case in Lemma 3.1 the cross-sectional space is not expanding, this argument does not immediately apply to the cylindrical case.
Up to this point, on we have constructed either a weight function on a cylindrical end, or a finite collection of weight functions on a subconic end, compatible on the end in the way described in Lemma 3.3. Now our next proposition provides the final modification of those weights to pull them back to .
Proposition 3.5 (Construction of Carleman weights).
On the prespecified , there are Carleman weights compatible with control from , in the sense of (2.1).
Proof.
1. We start by reviewing what we have learnt from the previous lemmata. Denote by , its closure by , and by as a matter of convenience. Lemma 3.1 and Lemma 3.3 state that we have an uniform with a finite family of weights on the ends , where on the end where it is defined, and specifically on some , with . Moreover there exists , such that for each , at , one has , and if for some we have then .
2. On , we start by constructing a weight on the central compactum, to which the ends on which we have the weights are attached. Set . On compact, there exists a Morse function with finitely many non-degenerate critical points, none of which resides on the boundary . Note that this can be achieved by finding a Morse function on a small closed neighbourhood of , for example , and find a diffeomorphism to move all critical points not on the new boundary into and then restrict the new function to . Apply a diffeomorphism on , to get where all critical points of are moved inside some given by the Network Control Condition. Note that we can assume without loss of generality that there is a such inside , by increasing if needed. Now construct
(3.33) |
Note that on , and for some positive away from . Now set
(3.34) |
We have on , and . These are weight functions on .
3. We now trim the parts of inside , where the supports of those weights could intersect, and extend them to the whole in a compatible manner. Construct two radial cutoff functions and in . Let be non-increasing, on , and on . Let be non-decreasing, on , and on , and on , and constant on . Meanwhile we ask on , which makes sense as has been taken large enough. See Figure LABEL:F2. Now set
(3.35) |
extended to the whole manifold by 0. Note are in .