This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exponential decay for damped Klein-Gordon equations on asymptotically cylindrical and conic manifolds

Ruoyu P. T. Wang Department of Mathematics, Northwestern University, 2033 Sheridan Road, Evanston, Illinois 60208, USA [email protected]
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 (M,g0)(M,g_{0}), a non-compact connected dd-dimensional manifold without boundary, with NN infinite ends,

M=M0¯(k=1NMk),M=\overline{M_{0}}\cup\left(\bigcup_{k=1}^{N}M_{k}\right), (1.1)

where M0¯\overline{M_{0}} is a compact, connected manifold with boundary M0¯=k=1N{1}×Mk\partial\overline{M_{0}}=\bigsqcup_{k=1}^{N}\{1\}\times\partial M_{k}. Denote the interior of M0¯\overline{M_{0}} by M0M_{0}. Each end MkM_{k} is identified as a cylinder (1,)r×Mk(1,\infty)_{r}\times\partial M_{k} endowed with a product metric

dr2+θ2k(r)hk,dr^{2}+\theta^{2}_{k}(r)h_{k}, (1.2)

where Mk\partial M_{k} is a (d1)(d-1)-dimensional compact manifold without boundary, hkh_{k} a smooth metric on Mk\partial M_{k}. The scaling functions θkC([1,);>0)\theta_{k}\in C^{\infty}([1,\infty);\mathbb{R}_{>0}) satisfy either one of the following conditions:

limrθk(r)=,|rmθk|Cm<,m1;or\displaystyle\lim_{r\rightarrow\infty}\theta_{k}(r)=\infty,\quad\lvert\partial_{r}^{m}\theta_{k}\rvert\leq C_{m}<\infty,\forall m\geq 1;\textrm{or} (1.3)
θk1.\displaystyle\theta_{k}\equiv 1. (1.4)

We call ends with the scaling functions θk\theta_{k} in (1.3) sub-conic ends, and ends with those in (1.4) cylindrical ends. Specifically, sub-conic ends with the scaling function θk(r)=r\theta_{k}(r)=r 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 (M,g)(M,g) that is a bounded perturbation of our model manifold (M,g0)(M,g_{0}), in the sense that both the identity map

Φ0:(M,g)(M,g0);pp\Phi_{0}:(M,g)\rightarrow(M,g_{0});\quad p\mapsto p (1.5)

and its inverse Φ01:(M,g0)(M,g)\Phi_{0}^{-1}:(M,g_{0})\rightarrow(M,g) are CbC^{\infty}_{b}-maps between two manifolds. See Appendix LABEL:SA for the definition of boundedness on manifolds of bounded geometry. Note that dΦ0(p)d\Phi_{0}(p) is uniformly bounded at each p(M,g)p\in(M,g), from both above and below as a map from TpMT_{p}M equipped with gg to TpMT_{p}M equipped with g0g_{0}.

We inexhaustively list some examples that are compatible with our setting:

Example 1.1.
  1. (1)

    Euclidean spaces d\mathbb{R}^{d}, with M0M_{0} being the unit open ball and M1M_{1} being the rest of d\mathbb{R}^{d} as [1,)r×𝕊d1[1,\infty)_{r}\times\mathbb{S}^{d-1} in spherical coordinates. This is a conic end with θ1(r)=r\theta_{1}(r)=r.

  2. (2)

    Euclidean spaces d\mathbb{R}^{d} as above, but endowed with a bounded perturbed metric, whose local matrix form in the canonical Euclidean coordinates, g(x)g(x), and its inverse g1(x)g^{-1}(x), are smooth matrix-valued functions of which components are Cb(d)C^{\infty}_{b}(\mathbb{R}^{d}).

  3. (3)

    Asymptotically conic manifolds, also as known as Riemannian scattering spaces, of finitely many ends of the form Mk=[0,1]x×MkM_{k}=[0,1]_{x}\times\partial M_{k} endowed with scattering metrics x4dx2+x2hkx^{-4}dx^{2}+x^{-2}h_{k}. Here hkh_{k}’s are smooth symmetric 2-cotensors on MkM_{k} whose restriction to Mk\partial M_{k} is positive-definite. In our model we realise the metric as a bounded perturbation of x4dx2+x2hk=dr2+r2hkx^{-4}dx^{2}+x^{-2}h_{k}^{\prime}=dr^{2}+r^{2}h_{k}^{\prime} where r=x1r=x^{-1} and hkh_{k}^{\prime}’s are metrics on Mk\partial M_{k} independent of rr. See [mel95] for further details.

  4. (4)

    Product cylinders of the form (,)×M(-\infty,\infty)\times\partial M where M\partial M is a closed manifold, by taking M0=(1,1)×MM_{0}=(-1,1)\times\partial M, [1,)×M1=M[1,\infty)\times M_{1}=\partial M and M2=(,1]×MM_{2}=(-\infty,-1]\times\partial M. 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 [0,1]x×Mk[0,1]_{x}\times\partial M_{k} endowed with x2dx2+hx^{-2}dx^{2}+h. Here again hh is a smooth symmetric 2-cotensor on MkM_{k} whose restriction to Mk\partial M_{k} is positive-definite. See [mel95] for further details.

  5. (5)

    Elliptic paraboloid, {(x,y,z):z=x2+y2}3\{(x,y,z):z=x^{2}+y^{2}\}\subset\mathbb{R}^{3} with M0M_{0} being the tip {z1}\{z\leq 1\} and M1M_{1} be the rest of paraboloid as [1,)r×𝕊1θ[1,\infty)_{r}\times\mathbb{S}^{1}_{\theta} equipped with metric (1+r1/4)dr2+rdθ2(1+r^{-1}/4)dr^{2}+rd\theta^{2}, under the change of coordinates (x,y,z)=(r1/2cosθ,r1/2sinθ,r)(x,y,z)=(r^{1/2}\cos{\theta},r^{1/2}\sin{\theta},r). Here this metric on M1M_{1} is a bounded perturbation of dr2+rdθ2dr^{2}+rd\theta^{2}, whose scaling function is θ1(r)=r1/2\theta_{1}(r)=r^{1/2}, which is sub-conic.

  6. (6)

    Boundedly perturbed cylinders. Consider the surface {(x,y,z):(2+cosz)2=x2+y2}3\{(x,y,z):(2+\cos{z})^{2}=x^{2}+y^{2}\}\subset\mathbb{R}^{3}, which is realised as the bounded perturbation to (,)r×𝕊1θ(-\infty,\infty)_{r}\times\mathbb{S}^{1}_{\theta} equipped with metric dr2+(2+cosr)dθ2dr^{2}+(2+\cos{r})d\theta^{2}. Note that, this surface is not an asymptotically cylindrical manifold, as the cosr\cos{r} cannot be well-defined at the spatial infinity r=r=\infty. But we could still cope with this manifold as a bounded perturbation of the product cylinder.

  7. (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 aCb(M)a\in C_{b}^{\infty}(M), a smooth function on MM whose derivatives of all orders are bounded by uniform constants dependent only on the order. The damped Klein-Gordon equation on our manifold (M,g)(M,g) reads

{(Δg+Id+t2+at)u(t,x)=0,on t0×Mxu(0,x)=u0(x)H2(M),tu(0,x)=u1(x)H1(M),\begin{cases}\left(\Delta_{g}+\operatorname{Id}+\partial_{t}^{2}+a\partial_{t}\right)u(t,x)=0,\quad\textrm{on }\mathbb{R}_{t\geq 0}\times M_{x}\\ u(0,x)=u_{0}(x)\in H^{2}(M),\quad\partial_{t}u(0,x)=u_{1}(x)\in H^{1}(M)\end{cases}, (1.6)

where Δg\Delta_{g} is the positive Laplace-Beltrami operator on (M,g)(M,g). Consider

A=(0Id(Δg+Id)a(x)),A=\begin{pmatrix}0&\operatorname{Id}\\ -\left(\Delta_{g}+\operatorname{Id}\right)&-a(x)\end{pmatrix}, (1.7)

which is a bounded linear operator from D(A)=H2(M)×H1(M)D(A)=H^{2}(M)\times H^{1}(M) to X=H1(M)×L2(M)X=H^{1}(M)\times L^{2}(M), which is further dissipative in the sense that

ReA(u,v),(u,v)X=Ma(x)|v(x)|2dg0.\operatorname{Re}{\left\langle A\left(u,v\right),(u,v)\right\rangle}_{X}=-\int_{M}a(x)\lvert v(x)\rvert^{2}~{}dg\leq 0. (1.8)

By noting that D(A)D(A) is dense in XX, AA is a bounded dissipative operator on Hilbert space XX, and the Lumer-Phillips theorem tells us AA generates a strongly continuous semigroup etAe^{tA} on XX that is further a contraction semigroup, in the sense that etAXX1\|e^{tA}\|_{X\rightarrow X}\leq 1 for each t0t\geq 0. Note we can formulate the equation (1.6) as a Cauchy problem for U(t,x)=(u(t,x),tu(t,x))U(t,x)=(u(t,x),\partial_{t}u(t,x)), that is

tU(t,x)=A(x)U(t,x),U(0,x)=U0(x)=(u0(x),u1(x))X,\partial_{t}U(t,x)=A(x)U(t,x),\quad U(0,x)=U_{0}(x)=(u_{0}(x),u_{1}(x))\in X, (1.9)

and the strongly continuous semigroup etAe^{tA} is the solution operator to the Cauchy problem, where the unique solution is etAU0e^{tA}U_{0}. As we look into how fast the global energy decays, it suffices to look at the decay of the operator norm of the semigroup etAe^{tA}. Indeed, the energy of the solution to (1.6) is

E(u,t)=12M|xu(t,x)|2+|tu(t,x)|2dx12etAXX(u0,u1)X.E(u,t)=\frac{1}{2}\int_{M}\lvert\nabla_{x}u(t,x)\rvert^{2}+\lvert\partial_{t}u(t,x)\rvert^{2}~{}dx\leq\frac{1}{2}\left\|e^{tA}\right\|_{X\rightarrow X}\left\|(u_{0},u_{1})\right\|_{X}. (1.10)

The semigroup etAe^{tA} weakly decays to 0 when the damping aa is smooth and not zero somewhere on MM, as in [wal77]. In this paper, we are interested in two types of decays of the semigroup etAe^{tA}: 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 MM 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 a(x)a(x), 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 d\mathbb{R}^{d}, 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 d\mathbb{R}^{d} under no further dynamical assumptions. In [jr18], the sharp polynomial global energy decay for the damped wave equation on d\mathbb{R}^{d} 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 (M,g)(M,g) 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 (M,g)(M,g):

Definition 1.2 (Geometric Control Condition).

We say the damping aa satisfies the Geometric Control Condition (T,α)(T,\alpha) on (M,g)(M,g), for T,α>0T,\alpha>0, if for (x,ξ)Σ(x,\xi)\in\Sigma, where Σ={(x,ξ)TM:|ξ|2=1}\Sigma=\{(x,\xi)\in T^{*}M:\lvert\xi\rvert^{2}=1\}, one has

aT(x,ξ)=1T0T((Πxφt)a)(x,ξ)dtα>0,\left<a\right>_{T}(x,\xi)=\frac{1}{T}\int_{0}^{T}\left(\left(\Pi_{x}\circ\varphi_{t}\right)^{*}a\right)(x,\xi)~{}dt\geq\alpha>0, (1.11)

where φt\varphi_{t} is the Hamiltonian flow associated with |ξ|g2\lvert\xi\rvert_{g}^{2}, and Πx\Pi_{x} is the projection from fibres of TMT^{*}M to the base variable.

We claim the first main result that Geometric Control Condition gives exponential decay of the semigroup etAe^{tA}:

Theorem 1 (Exponential decay of energy).

Assume aCb(M)a\in C^{\infty}_{b}(M) where a0a\geq 0 everywhere, satisfies the Geometric Control Condition (T,α)(T,\alpha), then the semigroup etAe^{tA} decays exponentially in the sense that etAXXMeλt\|e^{tA}\|_{X\rightarrow X}\leq Me^{-\lambda t}, for each t0t\geq 0, for some M,λ>0M,\lambda>0. It is then implied that the solution uu to the damped Klein-Gordon equation with initial datum (u0,u1)H2(M)×H1(M)(u_{0},u_{1})\in H^{2}(M)\times H^{1}(M),

{(Δg+Id+t2+at)u(t,x)=0,on t0×Mxu(0,x)=u0(x)H2(M),tu(0,x)=u1(x)H1(M).\begin{cases}\left(\Delta_{g}+\operatorname{Id}+\partial_{t}^{2}+a\partial_{t}\right)u(t,x)=0,\quad\textrm{on }\mathbb{R}_{t\geq 0}\times M_{x}\\ u(0,x)=u_{0}(x)\in H^{2}(M),\quad\partial_{t}u(0,x)=u_{1}(x)\in H^{1}(M)\end{cases}. (1.12)

decays exponentially, in the sense that there exists C,λ>0C,\lambda>0,

E(u,t)12Ceλt(u0H1(M)2+u1L2(M)2)1/2.E(u,t)\leq\frac{1}{2}Ce^{-\lambda t}\left(\left\|u_{0}\right\|_{H^{1}(M)}^{2}+\left\|u_{1}\right\|_{L^{2}(M)}^{2}\right)^{1/2}. (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 aa satisfies the Network Control Condition (L,ω,2β,{xn})(L,\omega,2\beta,\{x_{n}\}) on MM, for L,ω,2β>0L,\omega,2\beta>0, and {xn}\{x_{n}\} a set of points on MM, if at each xMx\in M,

d(x,n{xn})L,d(x,\bigcup_{n}\left\{x_{n}\right\})\leq L, (1.14)

and a(x)2β>0a(x)\geq 2\beta>0 on nB(xn,ω)\bigcup_{n}B(x_{n},\omega).

This hypothesis has been introduced in [bj16], and gives logarithmic decay on d\mathbb{R}^{d}. 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 aCb(M)a\in C^{\infty}_{b}(M) where a0a\geq 0 everywhere, satisfies the Network Control Condition (L,r,2β,{xn})(L,r,2\beta,\{x_{n}\}), then for each k1k\geq 1, the solution uu to the damped Klein-Gordon equation with initial datum (u0,u1)Hk+1(M)×Hk(M)(u_{0},u_{1})\in H^{k+1}(M)\times H^{k}(M),

{(Δg+Id+t2+at)u=0,on t0×Mxu|t=0=u0Hk+1(M),tu|t=0=u1Hk(M)\begin{cases}\left(\Delta_{g}+\operatorname{Id}+\partial_{t}^{2}+a\partial_{t}\right)u=0,\quad\textrm{on }\mathbb{R}_{t\geq 0}\times M_{x}\\ u|_{t=0}=u_{0}\in H^{k+1}(M),\quad\partial_{t}u|_{t=0}=u_{1}\in H^{k}(M)\end{cases} (1.15)

decays logarithmically, in the sense that there exists Ck>0C_{k}>0,

E(u)=(gu(t)2+tu(t)2)12Cklog(2+t)k(u0,u1)Hk+1×Hk.E(u)=\left(\left\|\nabla_{g}u(t)\right\|^{2}+\left\|\partial_{t}u(t)\right\|^{2}\right)^{\frac{1}{2}}\leq\frac{C_{k}}{\log\left(2+t\right)^{k}}\left\|\left(u_{0},u_{1}\right)\right\|_{H^{k+1}\times H^{k}}. (1.16)

Though the idea of the proof is similar to that of [bj16], we need new tools because of we are leaving d\mathbb{R}^{d}. In [bj16], they used the fact that i=1dcos(πxi)Cb(d)\prod_{i=1}^{d}\cos(\pi x_{i})\in C^{\infty}_{b}(\mathbb{R}^{d}) has critical points exactly at dd\mathbb{Z}^{d}\subset\mathbb{R}^{d}. On our (M,g)(M,g), neither the function nor the d\mathbb{Z}^{d}-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 aa in these two theorems can be weakened. In Theorem 1 we only need aL(M)a\in L^{\infty}(M) to be uniformly continuous, and in Theorem 2 we only need aL(M)a\in L^{\infty}(M). 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 (M,g)(M,g) 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 MM be a manifold of bounded geometry, without boundary. See Appendix LABEL:SA for further details. Let ΩM\Omega\subset M be an open set.

Definition 2.1 (Compatibility conditions).

We say a finite family of weight functions {ψ1,,ψn}Cb(M)\{\psi_{1},\dots,\psi_{n}\}\subset C^{\infty}_{b}(M) is compatible with control from Ω\Omega, if there exists an open set Ω0Ω\Omega_{0}\subset\Omega with the following properties:

  1. (1)

    We have d(Ω0,MΩ)>0d(\Omega_{0},M\setminus\Omega)>0 where dd is the distance on MM.

  2. (2)

    There exist constants ρ,τ>0\rho,\tau>0 such that, at each point xMΩ0x\in M\setminus\Omega_{0}, for each kk, if |gψk(x)|<2ρ\lvert\nabla_{g}\psi_{k}(x)\rvert<2\rho, then there exists some ll that

    |gψl(x)|2ρ,ψl(x)ψk(x)+τ.\lvert\nabla_{g}\psi_{l}(x)\rvert\geq 2\rho,\quad\psi_{l}(x)\geq\psi_{k}(x)+\tau. (2.1)

It is natural to impose the compatibility condition upon the weight functions. We aim to control the L2L^{2}-size of a quasi-mode by merely the L2L^{2}-size of that inside the region of control Ω\Omega. At xx 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 xx. Therefore there has to be another weight whose gradient is sufficiently large to control that locally near xx. This explains the first part of (2.1).

On another hand, at such a point xx, because we use the exponential weights exp(eλψl/h)\exp(e^{\lambda\psi_{l}}/h) 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 exp(eλψk/h)\exp(e^{\lambda\psi_{k}}/h). 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 exp(eλψk/h)eϵ/hexp(eλψl/h)\exp(e^{\lambda\psi_{k}}/h)\leq e^{-\epsilon/h}\exp(e^{\lambda\psi_{l}}/h) for some uniform constant ϵ>0\epsilon>0 depending on τ\tau. Now we note that the control induced by ψl\psi_{l} is observable, in the sense that the non-controlling weight ψk\psi_{k} generates an exponentially weaker term.

Theorem 3 (Global Carleman estimates with multiple weights).

For MM a manifold of bounded geometry without boundary, assume there are non-negative Carleman weights ψ1,,ψn\psi_{1},\dots,\psi_{n} compatible with the control from (Ω,Ω0)(\Omega,\Omega_{0}) in the sense of (2.1). Then, we have a global Carleman estimate with constant C>0C>0, independent of semiclassical parameter h(0,h0)h\in(0,h_{0}) for small h0h_{0}, such that

uL2(M)eC/h((h2ΔgV(x;h))uL2(M)+uL2(Ω)),\left\|u\right\|_{L^{2}(M)}\leq e^{C/h}\left(\left\|\left(h^{2}\Delta_{g}-V(x;h)\right)u\right\|_{L^{2}(M)}+\left\|u\right\|_{L^{2}(\Omega)}\right), (2.2)

where VCb(M×[0,h0])V\in C^{\infty}_{b}(M\times[0,h_{0}]) 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

V(x;h)=V0(x)+hV1(x)+h2V2(x;h),V(x;h)=V_{0}(x)+hV_{1}(x)+h^{2}V_{2}(x;h), (2.3)

where V0,V1Cb(M)V_{0},V_{1}\in C^{\infty}_{b}(M) and V2Cb(M×[0,h0])V_{2}\in C^{\infty}_{b}(M\times[0,h_{0}]). Fix a cutoff χCb(M)\chi\in C^{\infty}_{b}(M) such that χ1\chi\equiv 1 on MΩM\setminus\Omega and identically 0 on Ω0\Omega_{0}. Fix a ψk\psi_{k} and denote Ukυ={xM:|gψk(x)|<υ}U_{k}^{\upsilon}=\{x\in M:\lvert\nabla_{g}\psi_{k}(x)\rvert<\upsilon\}. For each kk, fix a χkCb(M)\chi_{k}\in C^{\infty}_{b}(M) such that χk1\chi_{k}\equiv 1 on MU2ρkM\setminus U^{2\rho}_{k} and identically 0 on UρkU^{\rho}_{k}. Set Ph=h2ΔgV(x;h)P_{h}=h^{2}\Delta_{g}-V(x;h). Construct the exponential Carleman weights by ϕk=eλψk\phi_{k}=e^{\lambda\psi_{k}}, where λ\lambda is some large number to be determined later, and the conjugated operator by

Pk,h=eϕk/hPheϕk/h=(h2Δg|gϕk|2V0(x)hV1(x)h2V2(x;h))+2hjϕkjhΔgϕk.P_{k,h}=e^{\phi_{k}/h}P_{h}e^{-\phi_{k}/h}=\left(h^{2}\Delta_{g}-\lvert\nabla_{g}\phi_{k}\rvert^{2}-V_{0}(x)-hV_{1}(x)-h^{2}V_{2}(x;h)\right)\\ +2h\nabla^{j}\phi_{k}\nabla_{j}-h\Delta_{g}\phi_{k}. (2.4)

See (LABEL:A1L8) for the notation jϕkj\nabla^{j}\phi_{k}\nabla_{j}. Note

Pk,hu2L2=Pk,hu2L2+[Pk,h,Pk,h]u,u[Pk,h,Pk,h]u,u=hOph(i1{pk,h¯,pk,h})+h2R2u,u=2hOph({pk,hR,pk,hI})+h2R2u,u,\left\|P_{k,h}u\right\|^{2}_{L^{2}}=\left\|P^{*}_{k,h}u\right\|^{2}_{L^{2}}+\left\langle[P_{k,h}^{*},P_{k,h}]u,u\right\rangle\geq\left\langle[P_{k,h}^{*},P_{k,h}]u,u\right\rangle\\ =h\left\langle\operatorname{Op}_{h}(i^{-1}\{\overline{p_{k,h}},p_{k,h}\})\right\rangle+h^{2}\langle R_{2}u,u\rangle=2h\left\langle\operatorname{Op}_{h}(\{p_{k,h}^{R},p_{k,h}^{I}\})\right\rangle+h^{2}\langle R_{2}u,u\rangle, (2.5)

where R2Ψu,h2R_{2}\in\Psi_{u,h}^{2} and the real and imaginary parts of the principal symbol are

pkR=|ξ|2|gϕk|2V0(x),pkI=2ξ(gϕk).p_{k}^{R}=\lvert\xi\rvert^{2}-\lvert\nabla_{g}\phi_{k}\rvert^{2}-V_{0}(x),\quad p_{k}^{I}=2\xi(\nabla_{g}\phi_{k}). (2.6)

Denote the subset of the cotangent bundle that contains the characteristic set,

Sk={(x,ξ)T(M(Ω0Ukρ)):14(|gϕk|2+V0(x))|ξ|24(|gϕk|2+V0(x))},S_{k}=\\ \left\{(x,\xi)\in T^{*}\left(M\setminus(\Omega_{0}\cup U_{k}^{\rho})\right):\frac{1}{4}\left(\lvert\nabla_{g}\phi_{k}\rvert^{2}+V_{0}(x)\right)\leq\lvert\xi\rvert^{2}\leq 4\left(\lvert\nabla_{g}\phi_{k}\rvert^{2}+V_{0}(x)\right)\right\}, (2.7)

outside of which (pkR)29/16(p_{k}^{R})^{2}\geq 9/16. Consider a microlocal cutoff bk(x,ξ)Su0(M)b_{k}(x,\xi)\in S_{u}^{0}(M) that is supported in SkS_{k} and is identically 1 on {(x,ξ)T(M(Ω0Uρk)):1/2(|gϕk|2+V0)1|ξ|22}\{(x,\xi)\in T^{*}(M\setminus(\Omega_{0}\cup U^{\rho}_{k})):1/2\leq(\lvert\nabla_{g}\phi_{k}\rvert^{2}+V_{0})^{-1}\lvert\xi\rvert^{2}\leq 2\}. Note Oph(1bk)\operatorname{Op}_{h}(1-b_{k}) and Oph((1bk)pk,h1)Pk,h\operatorname{Op}_{h}((1-b_{k})p_{k,h}^{-1})P_{k,h} have the same principal symbol, so

Oph(1bk)uH2hCPk,hu+ChuH1h.\left\|\operatorname{Op}_{h}(1-b_{k})u\right\|_{H^{2}_{h}}\leq C\left\|P_{k,h}u\right\|+Ch\left\|u\right\|_{H^{1}_{h}}. (2.8)

On another hand, let bk=bkξ2b_{k}^{\prime}=b_{k}\langle\xi\rangle^{2} then Oph(bk)\operatorname{Op}_{h}(b_{k}^{\prime}) and hD2Ophbk\langle hD\rangle^{2}\operatorname{Op}_{h}{b_{k}} are both in Ψu\Psi_{u}^{-\infty} and their principal symbols agree. Thus

Oph(bk)uH2hCOph(bk)u+ChuL2.\left\|\operatorname{Op}_{h}(b_{k})u\right\|_{H^{2}_{h}}\leq C\|\operatorname{Op}_{h}(b_{k}^{\prime})u\|+Ch\|u\|_{L^{2}}. (2.9)

From (2.8) and (2.9) we know that

uH2hCPk,hu+COph(bk)u.\left\|u\right\|_{H^{2}_{h}}\leq C\left\|P_{k,h}u\right\|+C\|\operatorname{Op}_{h}(b_{k}^{\prime})u\|. (2.10)

We claim that {pkR,pkI}\left\{p_{k}^{R},p_{k}^{I}\right\} is uniformly bounded from below on SkS_{k}. Note that at any (x,ξ)SkT(MUkρ)(x,\xi)\in S_{k}\subset T^{*}(M\setminus U_{k}^{\rho}) we have |gψk|ρ\lvert\nabla_{g}\psi_{k}\rvert\geq\rho from the definition of UkρU_{k}^{\rho}. Hence on SkS_{k} we have |gϕk|=λ|gψk|eλψkλρeλψk\lvert\nabla_{g}\phi_{k}\rvert=\lambda\lvert\nabla_{g}\psi_{k}\rvert e^{\lambda\psi_{k}}\geq\lambda\rho e^{\lambda\psi_{k}} and therefore

14(λ2ρ2e2λψk+V0(x))|ξ|24(λ2Cke2λψk+V0(x)),\frac{1}{4}\left(\lambda^{2}\rho^{2}e^{2\lambda\psi_{k}}+V_{0}(x)\right)\leq\lvert\xi\rvert^{2}\leq 4\left(\lambda^{2}C_{k}e^{2\lambda\psi_{k}}+V_{0}(x)\right), (2.11)

where CkC_{k}’s are some constants dependent only on the maximal size of first derivatives of ψk\psi_{k}’s. Let (x,ξ)Sk(x,\xi)\in S_{k}. In the canonical coordinates induced by the geodesic normal coordinates around xx, we have gij(x)=δijg_{ij}(x)=\delta_{ij} and gij(x)=0\nabla g_{ij}(x)=0. Compute, by noting that at xx we have g=\nabla_{g}=\nabla, the Euclidean gradient,

ξpkR.xpkI=4λ2eλψk|ψk|2|ξ|2+4λeλψkξt.(2ψk).ξ0+𝒪(λ3e3λψk)+𝒪(λeλψk)\partial_{\xi}p_{k}^{R}.\partial_{x}p_{k}^{I}=4\lambda^{2}e^{\lambda\psi_{k}}\lvert\nabla\psi_{k}\rvert^{2}\lvert\xi\rvert^{2}+4\lambda e^{\lambda\psi_{k}}\xi^{t}.(\nabla^{2}\psi_{k}).\xi\\ \geq 0+\mathcal{O}\left(\lambda^{3}e^{3\lambda\psi_{k}}\right)+\mathcal{O}\left(\lambda e^{\lambda\psi_{k}}\right) (2.12)

and

xpkR.ξpkI=2λ4e3λψk|ψk|4+2λeλψkψt.V0+λ3e3λψk(|ψk|2)t.ψk2ρ4λ4e3λψk+𝒪(λeλψk)+𝒪(λ3e3λψk),-\partial_{x}p_{k}^{R}.\partial_{\xi}p_{k}^{I}=2\lambda^{4}e^{3\lambda\psi_{k}}\lvert\nabla\psi_{k}\rvert^{4}+2\lambda e^{\lambda\psi_{k}}\nabla\psi^{t}.\nabla V_{0}\\ +\lambda^{3}e^{3\lambda\psi_{k}}\left(\nabla\lvert\nabla\psi_{k}\rvert^{2}\right)^{t}.\nabla\psi_{k}\geq 2\rho^{4}\lambda^{4}e^{3\lambda\psi_{k}}+\mathcal{O}\left(\lambda e^{\lambda\psi_{k}}\right)+\mathcal{O}\left(\lambda^{3}e^{3\lambda\psi_{k}}\right), (2.13)

from the uniform boundedness of the derivatives of order up to 2 of ψk\psi_{k}. Hence on SkS_{k} with a large λ\lambda,

{pkR,pkI}2ρ4λ4e3λψk+𝒪(λ3e3λψk)Cλ3e3λψk9/16>0.\left\{p_{k}^{R},p_{k}^{I}\right\}\geq 2\rho^{4}\lambda^{4}e^{3\lambda\psi_{k}}+\mathcal{O}\left(\lambda^{3}e^{3\lambda\psi_{k}}\right)\geq C\lambda^{3}e^{3\lambda\psi_{k}}\geq 9/16>0. (2.14)

We conclude that throughout T(M(Ω2βUkρ))T^{*}(M\setminus(\Omega_{2\beta}\cup U_{k}^{\rho})), we have

η(1bk)2ξ3+{pkR,pkI}9/16>0\eta(1-b_{k})^{2}\left\langle\xi\right\rangle^{3}+\left\{p_{k}^{R},p_{k}^{I}\right\}\geq 9/16>0 (2.15)

for some fixed η\eta large, and hence there exists C>0C>0 such that

η(1bk)2ξ3+{pkR,pkI}Cξ3\eta(1-b_{k})^{2}\left\langle\xi\right\rangle^{3}+\left\{p_{k}^{R},p_{k}^{I}\right\}\geq C\langle\xi\rangle^{3} (2.16)

Now by invoking the weak Garding inequality in Proposition LABEL:A1T1 on M(Ω0Ukρ)M\setminus(\Omega_{0}\cup U_{k}^{\rho}) we have for any uL2(M)u\in L^{2}(M) with support inside M(Ω2βUkρ)M\setminus(\Omega_{2\beta}\cup U_{k}^{\rho}),

Oph({pkR,pkI}+η(1bk)2ξ3)u,uCuH3/2h2.\left\langle\operatorname{Op}_{h}\left(\{p_{k}^{R},p_{k}^{I}\}+\eta(1-b_{k})^{2}\langle\xi\rangle^{3}\right)u,u\right\rangle\geq C\|u\|_{H^{3/2}_{h}}^{2}. (2.17)

This implies

Oph({pRk,pIk})u,uCuH3/2h2COph(1bk)uH3/2h2ChuH2h2.\left\langle\operatorname{Op}_{h}(\{p^{R}_{k},p^{I}_{k}\})u,u\right\rangle\geq C\|u\|_{H^{3/2}_{h}}^{2}-C\|\operatorname{Op}_{h}(1-b_{k})u\|_{H^{3/2}_{h}}^{2}-Ch\|u\|_{H^{2}_{h}}^{2}. (2.18)

From (2.5), we use (2.8) and (2.10) to obtain

Pk,hu2ChuH3/2h2ChPk,hu2\left\|P_{k,h}u\right\|^{2}\geq Ch\|u\|_{H^{3/2}_{h}}^{2}-Ch\|P_{k,h}u\|^{2} (2.19)

and absorb the last term: for any uu with support inside M(Ω2βUρk)M\setminus\left(\Omega_{2\beta}\cup U^{\rho}_{k}\right) we have

Pk,huCh12uH3/2h.\left\|P_{k,h}u\right\|\geq Ch^{\frac{1}{2}}\|u\|_{H^{3/2}_{h}}. (2.20)

Apply the above estimate to eϕk/hχkχue^{\phi_{k}/h}\chi_{k}\chi u to obtain the local estimate,

eϕk/hχkχuL2Ch12Pk,heϕk/hχkχuL2=Ch12eϕk/hPhχkχuL2\left\|e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\left\|P_{k,h}e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}=Ch^{-\frac{1}{2}}\left\|e^{\phi_{k}/h}P_{h}\chi_{k}\chi u\right\|_{L^{2}} (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

eϕk/hχkχuL2Ch12(eϕk/hχkχPhuL2+eϕk/h[Pk,h,χkχ]uL2)\left\|e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\left(\left\|e^{\phi_{k}/h}\chi_{k}\chi P_{h}u\right\|_{L^{2}}+\left\|e^{\phi_{k}/h}\left[P_{k,h},\chi_{k}\chi\right]u\right\|_{L^{2}}\right) (2.22)

and

[Ph,χkχ]u=h2g(χkgχ+χgχk)u2h2χkjχju2h2χjχkju.\left[P_{h},\chi_{k}\chi\right]u=h^{2}\nabla_{g}^{*}\left(\chi_{k}\nabla_{g}\chi+\chi\nabla_{g}\chi_{k}\right)u-2h^{2}\chi_{k}\nabla^{j}\chi\nabla_{j}u-2h^{2}\chi\nabla^{j}\chi_{k}\nabla_{j}u. (2.23)

Let κCb(M)\kappa\in C^{\infty}_{b}(M) be a cutoff function supported inside ΩΩ0\Omega\setminus\Omega_{0} and being identically 11 on the support of gχ\nabla_{g}\chi, and let κkCb(M)\kappa_{k}\in C^{\infty}_{b}(M) be cutoff functions supported inside U2ρkUρkU^{2\rho}_{k}\setminus U^{\rho}_{k} and identically 11 on the support of gχk\nabla_{g}\chi_{k}. We immediately have the estimates of the first two terms in (2.23),

h2g(χkgχ+χgχk)uL2Ch2κuL2+Ch2κkuL2,\left\|h^{2}\nabla_{g}^{*}\left(\chi_{k}\nabla_{g}\chi+\chi\nabla_{g}\chi_{k}\right)u\right\|_{L^{2}}\leq Ch^{2}\left\|\kappa u\right\|_{L^{2}}+Ch^{2}\left\|\kappa_{k}u\right\|_{L^{2}}, (2.24)

from the uniform boundedness of first two derivatives of χ\chi and χk\chi_{k}. Consider the next term,

2h2χkjχjuL224h4|χkgχ||gu|2L2=4h4|χkgχ|2gu,gu=4h4g(|χkgχ|2gu),u=4h4Rej|χkgχ|2ju,u+4h4Re|χkgχ|2Δgu,u.\left\|2h^{2}\chi_{k}\nabla^{j}\chi\nabla_{j}u\right\|_{L^{2}}^{2}\leq 4h^{4}\left\|\lvert\chi_{k}\nabla_{g}\chi\rvert\lvert\nabla_{g}u\rvert\right\|^{2}_{L^{2}}=4h^{4}\left\langle\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\nabla_{g}u,\nabla_{g}u\right\rangle\\ =4h^{4}\left\langle\nabla^{*}_{g}\left(\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\nabla_{g}u\right),u\right\rangle=-4h^{4}\operatorname{Re}\left\langle\nabla^{j}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\nabla_{j}u,u\right\rangle\\ +4h^{4}\operatorname{Re}\left\langle\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\Delta_{g}u,u\right\rangle. (2.25)

The first term of the last line is estimated via an adjoint argument

4h4Re(j|χkgχ|2)ju,u=2h4(Δg|χkgχ|2)u,u.4h^{4}\operatorname{Re}\left\langle\left(\nabla^{j}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)\nabla_{j}u,u\right\rangle=2h^{4}\left\langle\left(\Delta_{g}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)u,u\right\rangle. (2.26)

Indeed, we have

(j|χkgχ|2)ju,u=gu,(g|χkgχ|2)u=u,g((g|χkgχ|2)u)=u,((Δg|χkgχ|2)u)u,(j|χkgχ|2)ju.\left\langle\left(\nabla^{j}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)\nabla_{j}u,u\right\rangle=\left\langle\nabla_{g}u,\left(\nabla_{g}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)u\right\rangle\\ =\left\langle u,\nabla_{g}^{*}\left(\left(\nabla_{g}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)u\right)\right\rangle=\left\langle u,\left(\left(\Delta_{g}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)u\right)\right\rangle-\left\langle u,\left(\nabla^{j}\lvert\chi_{k}\nabla_{g}\chi\rvert^{2}\right)\nabla_{j}u\right\rangle. (2.27)

We bring (2.25) and (2.26) together to see

2h2χkjχjuL2Ch2κuL2+Ch2κΔguL2=Ch2κuL2+CκPhuL2+CκV(x)uL2CκuL2+CPhuL2,\left\|2h^{2}\chi_{k}\nabla^{j}\chi\nabla_{j}u\right\|_{L^{2}}\leq Ch^{2}\left\|\kappa u\right\|_{L^{2}}+Ch^{2}\left\|\kappa\Delta_{g}u\right\|_{L^{2}}=Ch^{2}\left\|\kappa u\right\|_{L^{2}}\\ +C\|\kappa P_{h}u\|_{L^{2}}+C\|\kappa V(x)u\|_{L^{2}}\leq C\left\|\kappa u\right\|_{L^{2}}+C\|P_{h}u\|_{L^{2}}, (2.28)

since the third order derivatives of χ\chi are uniformly bounded and V(x)V(x) is a bounded potential. Symmetrically we have

2h2χjχkjuL2CκkuL2+CPhuL2\left\|2h^{2}\chi\nabla^{j}\chi_{k}\nabla_{j}u\right\|_{L^{2}}\leq C\left\|\kappa_{k}u\right\|_{L^{2}}+C\|P_{h}u\|_{L^{2}} (2.29)

and a complete estimate of (2.23),

[Ph,χkχ]uL2CPhuL2+CκuL2+CκkuL2.\left\|\left[P_{h},\chi_{k}\chi\right]u\right\|_{L^{2}}\leq C\|P_{h}u\|_{L^{2}}+C\left\|\kappa u\right\|_{L^{2}}+C\left\|\kappa_{k}u\right\|_{L^{2}}. (2.30)

Finally we have for h(0,h0]h\in(0,h_{0}],

eϕk/hχkχuL2Ch12eϕk/hPhuL2+Ch12eϕk/hκkuL2+Ch12eϕk/hκuL2.\left\|e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\left\|e^{\phi_{k}/h}P_{h}u\right\|_{L^{2}}+Ch^{-\frac{1}{2}}\left\|e^{\phi_{k}/h}\kappa_{k}u\right\|_{L^{2}}\\ +Ch^{-\frac{1}{2}}\left\|e^{\phi_{k}/h}\kappa u\right\|_{L^{2}}. (2.31)

Note that the constants λ,C,h0>0\lambda,C,h_{0}>0 could be chosen uniformly such that (2.31) holds for each kk. Sum up (2.31) over k=1,,nk=1,\dots,n to get the crude version of the global estimate,

k=1neϕk/hχkχuL2Ch12k=1neϕk/hPhuL2+Ch12k=1neϕk/hκkuL2+Ch12k=1neϕk/hκuL2,\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}P_{h}u\right\|_{L^{2}}+Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}\kappa_{k}u\right\|_{L^{2}}\\ +Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}\kappa u\right\|_{L^{2}}, (2.32)

in which κ\kappa is supported inside ΩΩ0\Omega\setminus\Omega_{0}, and κk\kappa_{k} is supported inside U2ρkUρkU^{2\rho}_{k}\setminus U^{\rho}_{k}.

3. Finally we use the compatibility condition (2.1) imposed upon the Carleman weights to refine the global estimate (2.32). Recall that Ul2ρ={xM:|gψl|<2ρ}U_{l}^{2\rho}=\{x\in M:\lvert\nabla_{g}\psi_{l}\rvert<2\rho\} stands for the points at which the weight ψl\psi_{l} fails to control the quasimode. Given the assumptions on compatibility (2.1), at each xUl2ρx\in U_{l}^{2\rho}, there exists some mm such that

ψlψmτ,ϕmϕl+(eλτ1)eλψlϕl+ϵ,\psi_{l}\leq\psi_{m}-\tau,\quad\phi_{m}\geq\phi_{l}+\left(e^{\lambda\tau}-1\right)e^{\lambda\psi_{l}}\geq\phi_{l}+\epsilon, (2.33)

where

ϵ=(eλτ1)eλmink=1n(infMψk)>0.\epsilon=\left(e^{\lambda\tau}-1\right)e^{\lambda\min_{k=1}^{n}(\inf_{M}\psi_{k})}>0. (2.34)

Note that ϵ\epsilon does not depend on l,ml,m. We have

eϕl/heϵ/heϕm/h.e^{\phi_{l}/h}\leq e^{-\epsilon/h}e^{\phi_{m}/h}. (2.35)

Now at each xMΩ0x\in M\setminus\Omega_{0}, we can partition {1,,n}\{1,\dots,n\} into {l1,,lnq}\{l_{1},\dots,l_{n-q}\} and {m1,,mq}\{m_{1},\dots,m_{q}\} with some 0<qn0<q\leq n, where

|gϕl(x)|<2ρ,|gϕm(x)|2ρ.\lvert\nabla_{g}\phi_{l_{*}}(x)\rvert<2\rho,\quad\lvert\nabla_{g}\phi_{m_{*}}(x)\rvert\geq 2\rho. (2.36)

For each i=1,,nqi=1,\dots,n-q, as xU2ρlix\in U^{2\rho}_{l_{i}}, there is some mjm_{j} such that

ϕliϕmjϵ,eϕli/heϵ/heϕmj/heϵ/hj=1qeϕmj/h,\phi_{l_{i}}\leq\phi_{m_{j}}-\epsilon,\quad e^{\phi_{l_{i}}/h}\leq e^{-\epsilon/h}e^{\phi_{m_{j}}/h}\leq e^{-\epsilon/h}\sum_{j=1}^{q}e^{\phi_{m_{j}}/h}, (2.37)

from (2.33) and (2.35). Then

i=1nqeϕli/h(nq)eϵ/hj=1qeϕmj/h.\sum_{i=1}^{n-q}e^{\phi_{l_{i}}/h}\leq\left(n-q\right)e^{-\epsilon/h}\sum_{j=1}^{q}e^{\phi_{m_{j}}/h}. (2.38)

As xUmj2ρx\notin U_{m_{j}}^{2\rho}, the cutoffs χmj\chi_{m_{j}}’s are all 11, and therefore

k=1neϕk/hχkqj=1eϕmj/h12qj=1eϕmj/h+12(nq)eϵ/hlinqeϕli/h12nk=1neϕk/h,\sum_{k=1}^{n}e^{\phi_{k}/h}\chi_{k}\geq\sum^{q}_{j=1}e^{\phi_{m_{j}}/h}\geq\frac{1}{2}\sum^{q}_{j=1}e^{\phi_{m_{j}}/h}+\frac{1}{2(n-q)}e^{\epsilon/h}\sum_{l_{i}}^{n-q}e^{\phi_{l_{i}}/h}\\ \geq\frac{1}{2n}\sum_{k=1}^{n}e^{\phi_{k}/h}, (2.39)

for 0<q<n0<q<n. Note that when q=nq=n, (2.39) holds trivially because all χk\chi_{k}’s are 1, and therefore we conclude that (2.39) holds at each xMΩ0x\in M\setminus\Omega_{0}. This improves the estimate (2.32) from the left:

k=1neϕk/hχkχuL2k=1neϕk/hχkχuL212n(k=1neϕk/h)χuL2\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\geq\left\|\sum_{k=1}^{n}e^{\phi_{k}/h}\chi_{k}\chi u\right\|_{L^{2}}\geq\frac{1}{2n}\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)\chi u\right\|_{L^{2}} (2.40)

as χ\chi is supported inside MΩ0M\setminus\Omega_{0}.

Meanwhile, for each l=1,,nl=1,\dots,n, the cutoff function κl\kappa_{l} is supported in Ul2ρU_{l}^{2\rho}. Therefore at each xUl2ρx\in U_{l}^{2\rho}, we know from (2.33) and (2.35) that

eϕl/hκleϵ/heϕm/heϵ/hk=1neϕk/h.e^{\phi_{l}/h}\kappa_{l}\leq e^{-\epsilon/h}e^{\phi_{m}/h}\leq e^{-\epsilon/h}\sum_{k=1}^{n}e^{\phi_{k}/h}. (2.41)

This inequality outside Ul2ρU_{l}^{2\rho} holds trivially as κl\kappa_{l} vanishes, and hence holds everywhere in MΩ0M\setminus\Omega_{0}. It improves the second term on the right in (2.32),

Ch12l=1neϕl/hκluL2Ch12l=1neϵ/hk=1neϕk/huL2=Cnh12eϵ/h(k=1neϕk/h)uL2.Ch^{-\frac{1}{2}}\sum_{l=1}^{n}\left\|e^{\phi_{l}/h}\kappa_{l}u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\sum_{l=1}^{n}\left\|e^{-\epsilon/h}\sum_{k=1}^{n}e^{\phi_{k}/h}u\right\|_{L^{2}}\\ =Cnh^{-\frac{1}{2}}e^{-\epsilon/h}\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)u\right\|_{L^{2}}. (2.42)

Bring (2.40) and (2.42) into (2.32) to observe

12n(k=1neϕk/h)χuL2Ch12k=1neϕk/hPhuL2+Cnh12eϵ/h(k=1neϕk/h)uL2+Ch12k=1neϕk/hκuL2.\frac{1}{2n}\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)\chi u\right\|_{L^{2}}\leq Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}P_{h}u\right\|_{L^{2}}\\ +Cnh^{-\frac{1}{2}}e^{-\epsilon/h}\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)u\right\|_{L^{2}}+Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}\kappa u\right\|_{L^{2}}. (2.43)

Finally, as 1χ1-\chi and κ\kappa are both supported inside Ω\Omega and are uniformly bounded, we can bound the L2(M)L^{2}(M)-norm of terms eϕk/h(1χ)ue^{\phi_{k}/h}(1-\chi)u and eϕk/hκue^{\phi_{k}/h}\kappa u by the L2(Ω)L^{2}(\Omega)-norm of eϕk/hue^{\phi_{k}/h}u, that is,

(k=1neϕk/h)uL2(M)(k=1neϕk/h)(1χ)uL2(M)+(k=1neϕk/h)χuL2(M)Ch12k=1neϕk/hPhuL2(M)+Cnh12eϵ/h(k=1neϕk/h)uL2(M)+Ch12k=1neϕk/huL2(Ω)\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)u\right\|_{L^{2}(M)}\leq\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)\left(1-\chi\right)u\right\|_{L^{2}(M)}\\ +\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)\chi u\right\|_{L^{2}(M)}\leq Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}P_{h}u\right\|_{L^{2}(M)}\\ +Cnh^{-\frac{1}{2}}e^{-\epsilon/h}\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)u\right\|_{L^{2}(M)}+Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}u\right\|_{L^{2}(\Omega)} (2.44)

As h12eϵ/h0h^{-\frac{1}{2}}e^{-\epsilon/h}\rightarrow 0 semiclassically, we can absorb the second term on the right by the term on the left for small hh, that is,

(k=1neϕk/h)uL2(M)Ch12k=1neϕk/hPhuL2(M)+Ch12k=1neϕk/huL2(Ω).\left\|\left(\sum_{k=1}^{n}e^{\phi_{k}/h}\right)u\right\|_{L^{2}(M)}\leq Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}P_{h}u\right\|_{L^{2}(M)}\\ +Ch^{-\frac{1}{2}}\sum_{k=1}^{n}\left\|e^{\phi_{k}/h}u\right\|_{L^{2}(\Omega)}. (2.45)

Denote the global maximum and minimum over all ϕk\phi_{k}’s by

K+=max1kn(supxMϕk(x)),K=min1kn(infxMϕk(x)),K_{+}=\max_{1\leq k\leq n}\left(\sup_{x\in M}\phi_{k}(x)\right),\quad K_{-}=\min_{1\leq k\leq n}\left(\inf_{x\in M}\phi_{k}(x)\right), (2.46)

where K+>KK_{+}>K_{-}. Then we have from (2.45),

eK/huL2(M)Cnh12eK+/hPhuL2(M)+Cnh12eK+/huL2(Ω),e^{{K_{-}}/h}\left\|u\right\|_{L^{2}(M)}\leq Cnh^{-\frac{1}{2}}e^{{K_{+}}/h}\left\|P_{h}u\right\|_{L^{2}(M)}+Cnh^{-\frac{1}{2}}e^{{K_{+}}/h}\left\|u\right\|_{L^{2}(\Omega)}, (2.47)

which is reduced to

uL2(M)eC/h(PhuL2(M)+uL2(Ω)).\left\|u\right\|_{L^{2}(M)}\leq e^{C/h}\left(\left\|P_{h}u\right\|_{L^{2}(M)}+\left\|u\right\|_{L^{2}(\Omega)}\right). (2.48)

This is our claim. ∎

Remark 2.2.
  1. (1)

    When MM is a compact manifold without boundary, then one could control from any open set Ω\Omega with arbitrary open subset Ω0\Omega_{0} with only one weight. It suffices to find a Morse function and find a diffeomorphism moving all critical points into Ω0\Omega_{0}.

  2. (2)

    It is observed that the uniform gap ψlψk+τ\psi_{l}\geq\psi_{k}+\tau in the compatibility condition is necessary. By this fixed gap τ\tau, we extracted an eε/he^{-\varepsilon/h}-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 L2(M)L^{2}(M)-norm of k=1neϕk/hu\sum_{k=1}^{n}e^{\phi_{k}/h}u, as this term on the right is now of size h12h^{-\frac{1}{2}}\rightarrow\infty semiclassically.

3. Construction of Carleman weights

In this section, we aim to explicitly construct the weight functions on our prespecified manifold (M,g)(M,g) in (1.5) to obtain the global Carleman estimate we developed in the previous section. Assume throughout this section that the Network Control Condition (L,ω,2β,{xm})(L,\omega,2\beta,\{x_{m}\}) defined in Definition 1.3 holds on (M,g)(M,g). Let Ωυ={xM:a(x)>υ}\Omega_{\upsilon}=\{x\in M:a(x)>\upsilon\}. Our ultimate target in this section is to control the whole manifold (M,g)(M,g) from (Ωβ,Ω2β)\left(\Omega_{\beta},\Omega_{2\beta}\right).

The strategy is to start by working on the model manifold (M,g0)(M,g_{0}) 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 (Ωβ,Ω2β)(\Omega_{\beta},\Omega_{2\beta}). Eventually we pull back the weights via Φ0\Phi_{0} back to the prespecified (M,g)(M,g).

Note that Φ0(x)=x\Phi_{0}(x)=x for each xMx\in M, and Φ0(Ωυ)=Ωυ\Phi_{0}(\Omega_{\upsilon})=\Omega_{\upsilon}. We claim that a Network Control Condition (L,ω,2β,{xm})(L,\omega,2\beta,\{x_{m}\}) on (M,g)(M,g) implies another (L,ω,2β,{xm})(L^{\prime},\omega^{\prime},2\beta,\{x_{m}\}) on (M,g0)(M,g_{0}). Recall that dΦ01d\Phi_{0}^{-1} is bounded from above and below,

C0dΦ01C1.C_{0}\leq\left\|d\Phi_{0}^{-1}\right\|\leq C_{1}. (3.1)

and therefore

C11dg(x,y)dg0(Φ0(x),Φ0(y))C01dg(x,y)C_{1}^{-1}d_{g}(x,y)\leq d_{g_{0}}(\Phi_{0}(x),\Phi_{0}(y))\leq C_{0}^{-1}d_{g}(x,y) (3.2)

for each x,yMx,y\in M. Let L=C01LL^{\prime}=C_{0}^{-1}L and ω=C01ω\omega^{\prime}=C_{0}^{-1}\omega. At each xMx\in M, we have

dg0(x,m{xm})C01dg(x,m{xm})C01L=L,d_{g_{0}}(x,\bigcup_{m}\left\{x_{m}\right\})\leq C_{0}^{-1}d_{g}(x,\bigcup_{m}\left\{x_{m}\right\})\leq C_{0}^{-1}L=L^{\prime}, (3.3)

the last inequality of which comes from the Network Control Condition (L,ω,2β,{xm})(L,\omega,2\beta,\{x_{m}\}) on (M,g)(M,g). We also have a(x)2β>0a(x)\geq 2\beta>0 on mBg0(xm,ω)mBg(xm,ω)\bigcup_{m}B_{g_{0}}(x_{m},\omega^{\prime})\subset\bigcup_{m}B_{g}(x_{m},\omega) as an immediate result of (3.2). Therefore we could, without loss of generality, assume a Network Control Condition (L,ω,2β,{xm})(L,\omega,2\beta,\{x_{m}\}) on (M,g0)(M,g_{0}). Note that

mBg0(xm,ω)Ω2β.\bigcup_{m}B_{g_{0}}(x_{m},\omega)\subset\Omega_{2\beta}. (3.4)

We begin by constructing weight functions on the cylindrical ends, where the scaling functions θk\theta_{k}’s are identically 1.

Lemma 3.1 (Cylindrical ends).

Consider a cylindrical end (Mk,g0)(M_{k},g_{0}), that is Mk×(1,)r\partial M_{k}\times(1,\infty)_{r} endowed with the metric g0=dr2+hg_{0}=dr^{2}+h, where hh is a smooth metric on closed Mk\partial M_{k}. There exists ψCb(Mk)\psi\in C^{\infty}_{b}(M_{k}), where 1ψ31\leq\psi\leq 3 and there is some ρ>0\rho>0 such that,

xΩ2β|g0ψ(x)|2ρ.x\notin\Omega_{2\beta}\Rightarrow\lvert\nabla_{g_{0}}\psi(x)\rvert\geq 2\rho. (3.5)
Proof.

In this lemma we Let M=MkM=M_{k} for we here only care what is happening on MkM_{k}.

1. We start by constructing a prototype weight based on the periodic structure. On M\partial{M} pick a Morse function ψ0C(M)\psi_{0}\in C^{\infty}(\partial M) that is positive on M\partial M. As M\partial M is compact, ψ0\psi_{0} has NN critical points at p1,,pNMp_{1},\dots,p_{N}\in\partial M. Fix ϵ>0\epsilon>0 small. Let a periodic function ψ1(r)C([1,))\psi_{1}(r)\in C^{\infty}(\left[1,\infty\right)) be given by

ψ1(r)=cos(π(r(1+2ϵ))2(L+4ω)(N+1))+2,\psi_{1}(r)=\cos\left(\frac{\pi\left(r-\left(1+2\epsilon\right)\right)}{2\left(L+4\omega\right)\left(N+1\right)}\right)+2, (3.6)

for small ϵ>0\epsilon>0. This is a function with a period of 4(L+4ω)(N+1)4\left(L+4\omega\right)\left(N+1\right). Consider

ψ~2(y,r)=ψ0(y)ψ1(r),\tilde{\psi}_{2}(y,r)=\psi_{0}(y)\psi_{1}(r), (3.7)

and modify its size to get

ψ2=1+2(maxψ~2minψ~2)1(ψ~2minψ~2),\psi_{2}=1+2\left(\max{\tilde{\psi}_{2}}-\min{\tilde{\psi}_{2}}\right)^{-1}\left(\tilde{\psi}_{2}-\min{\tilde{\psi}_{2}}\right), (3.8)

where we note that 1ψ231\leq\psi_{2}\leq 3 and we will later modify ψ2\psi_{2} to move around the critical points. The critical points of ψ2\psi_{2} are

pk,t=(pk,1+2ϵ+2t(L+4r)(N+1)),t0,k=1,,N,p_{k,t}=\left(p_{k},1+2\epsilon+2t\left(L+4r\right)\left(N+1\right)\right),\quad t\in\mathbb{N}_{0},\quad k=1,\dots,N, (3.9)

where 0={0}\mathbb{N}_{0}=\mathbb{N}\cup\{0\}.

M\partial Mp1,0p_{1,0}p1,0p_{1,0}^{\prime}p2,0p_{2,0}p2,0p_{2,0}^{\prime}p3,0p_{3,0}p3,0p_{3,0}^{\prime}p4,0p_{4,0}p4,0p_{4,0}^{\prime}
Figure 1. Φ2\Phi_{2} stretches the critical points of ψ2\psi_{2} apart.

2. We modify the weight to have critical points of distance uniformly bounded from below by 2(L+4ω)2(L+4\omega) from each other. For 1kN1\leq k\leq N, define the flows γk(s)\gamma_{k}(s) for s[0,)s\in[0,\infty) by

γk(s):(y,r)(y,r+2k(L+4ω)s),\gamma_{k}(s):\left(y,r\right)\mapsto\left(y,r+2k\left(L+4\omega\right)s\right), (3.10)

generated by the constant radial vector fields 2k(L+4ω)r2k\left(L+4\omega\right)\partial_{r}. Also denote by the flow γ(t)\gamma(t),

γ(t):(y,r)(y,r+4(N+1)(L+4ω)t),t0\gamma(t):\left(y,r\right)\mapsto\left(y,r+4(N+1)\left(L+4\omega\right)t\right),\quad t\in\mathbb{N}_{0} (3.11)

that preserves the periodicity of ψ2\psi_{2}, in the sense that γ(t)ψ2=ψ2\gamma(t)^{*}\psi_{2}=\psi_{2} for each t0t\in\mathbb{N}_{0}. Note that the flow γk(1)\gamma_{k}(1) pulls back points

pk,t=(pk,1+2ϵ+2t(L+4r)(N+1)+2k(L+4ω)),t0p^{\prime}_{k,t}=\left(p_{k},1+2\epsilon+2t\left(L+4r\right)\left(N+1\right)+2k\left(L+4\omega\right)\right),\quad t\in\mathbb{N}_{0} (3.12)

to critical points pk,tp_{k,t}, that is, γk(1):pk,tpk,t\gamma_{k}(1):p_{k,t}\mapsto p^{\prime}_{k,t}. Let Γk,t=γk([0,1])(pk,t)\Gamma_{k,t}=\gamma_{k}\left([0,1]\right)(p_{k,t}) and Γk,tυ\Gamma_{k,t}^{\upsilon} to be the υ\upsilon-neighbourhood of Γk,t\Gamma_{k,t}, for υ=ϵ,ϵ/2\upsilon=\epsilon,\epsilon/2. Now for t=0,1t=0,1, construct diffeomorphisms ϕk,tC(Γk,tϵ;Γk,tϵ)\phi_{k,t}\in C^{\infty}(\Gamma_{k,t}^{\epsilon};\Gamma_{k,t}^{\epsilon}) with inverses ϕ1k,tC(Γk,tϵ;Γk,tϵ)\phi^{-1}_{k,t}\in C^{\infty}(\Gamma_{k,t}^{\epsilon};\Gamma_{k,t}^{\epsilon}) such that

{ϕk,t:pk,t=γk(1)pk,tpk,tϕk,t=Id,onΓk,tϵΓk,tϵ/2.\begin{cases}\phi_{k,t}:p^{\prime}_{k,t}=\gamma_{k}(1)p_{k,t}\mapsto p_{k,t}\\ \phi_{k,t}=\operatorname{Id},~{}\textrm{on}~{}{\Gamma_{k,t}^{\epsilon}\setminus\Gamma_{k,t}^{\epsilon/2}}\end{cases}. (3.13)

Because that all Γk,tϵ\Gamma_{k,t}^{\epsilon}’s are disjoint and ϕk,t=Id\phi_{k,t}=\operatorname{Id} on Γk,tϵΓk,tϵ/2\Gamma_{k,t}^{\epsilon}\setminus\Gamma_{k,t}^{\epsilon/2}, we can glue up ϕk,t\phi_{k,t}’s to obtain a diffeomorphism Φ1\Phi_{1} on M×[1,1+4(L+4r)(N+1)]\partial M\times\left[1,1+4\left(L+4r\right)\left(N+1\right)\right]. Note that Φ1,Φ11\Phi_{1},\Phi_{1}^{-1} have all derivatives uniformly bounded from above and below, as the domain is compact. Also note that Φ1\Phi_{1} is the identity on M×[1,1+ϵ]\partial M\times[1,1+\epsilon] and M×[1+4(L+4r)(N+1)ϵ,1+4(L+4r)(N+1)]\partial M\times[1+4\left(L+4r\right)\left(N+1\right)-\epsilon,1+4\left(L+4r\right)\left(N+1\right)]. This enables us to extend Φ1\Phi_{1} periodically to some Φ2\Phi_{2} on MM, by defining on M×[1+4t(L+4r)(N+1),1+4(t+1)4t(L+4r)(N+1)]\partial M\times[1+4t(L+4r)(N+1),1+4(t+1)4t(L+4r)(N+1)], for each t0t\in\mathbb{N}_{0},

Φ2=γ1(t)Φ1γ(t).\Phi_{2}=\gamma^{-1}(t)^{*}\Phi_{1}\gamma(t)^{*}. (3.14)

Let ψ3=Φ2ψ2\psi_{3}=\Phi_{2}^{*}\psi_{2}, whose critical points are

pk,t=(pk,1+2ϵ+2t(L+4ω)(N+1)+2k(L+4ω)).p^{\prime}_{k,t}=\left(p_{k},1+2\epsilon+2t\left(L+4\omega\right)\left(N+1\right)+2k\left(L+4\omega\right)\right). (3.15)

for each t0t\in\mathbb{N}_{0}, each k=1,,Nk=1,\dots,N. Renumber those critical points by pmp^{\prime}_{m}. We remark that any two critical points of ψ3{\psi_{3}} are separated by distance of at least 2(L+4ω)2\left(L+4\omega\right). We also note for each R>0R>0 one has |g0ψ3|C\lvert\nabla_{g_{0}}\psi_{3}\rvert\geq C outside mB(pm,R)\bigcup_{m}B(p^{\prime}_{m},R) for constant C>0C>0 only depending on RR, because ψ3\psi_{3} 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

B¯(pm,L+3ω)B¯(pm,L+3ω)=\bar{B}\left(p^{\prime}_{m},L+3\omega\right)\cap\bar{B}\left(p^{\prime}_{m^{\prime}},L+3\omega\right)=\emptyset (3.16)

for any pair of critical points pmp^{\prime}_{m} and pmp^{\prime}_{m^{\prime}}. By the Network Control Condition, in each ball B(pm,L+2ω)B(p^{\prime}_{m},L+2\omega) we can find some xmx_{m} in the network such that

B¯(xm,ω)B(pm,L+2ω),\bar{B}\left(x_{m},\omega\right)\subset B\left(p^{\prime}_{m},L+2\omega\right), (3.17)

and a2βa\geq 2\beta on B¯(xm,ω)\bar{B}(x_{m},\omega). Now in each ball B(pm,L+3ω)B(p^{\prime}_{m},L+3\omega), find a diffeomorphism ϕm\phi^{\prime}_{m} such that

{ϕm:xmpmϕm=Id,onB(pm,L+3ω)B(pm,L+2ω).\begin{cases}\phi_{m}^{\prime}:x_{m}\mapsto p^{\prime}_{m}\\ \phi_{m}^{\prime}=\operatorname{Id},~{}\textrm{on}~{}B\left(p^{\prime}_{m},L+3\omega\right)\setminus B\left(p^{\prime}_{m},L+2\omega\right)\end{cases}. (3.18)

Glue up ϕm\phi^{\prime}_{m}’s to get a diffeomorphism Φ3\Phi_{3} on MM. We remark here that we can make this construction uniform in the sense that both Φ3,Φ31\Phi_{3},\Phi_{3}^{-1} are in Cb(M)C_{b}^{\infty}(M), as in [bj16, rm16]. Therefore we have B(pm,R)=B(Φ3(xm),R)Φ3(B(xm,ω))B(p^{\prime}_{m},R)=B(\Phi_{3}\left(x_{m}\right),R)\subset\Phi_{3}\left(B\left(x_{m},\omega\right)\right) for some R>0R>0 uniform in all mm. Now set ψ=Φ3ψ3\psi=\Phi^{*}_{3}{\psi_{3}}. We know |g0ψ3|C>0\lvert\nabla_{g_{0}}{\psi_{3}}\rvert\geq C>0 uniformly for all pmB(pm,R)p\notin\cup_{m}B(p^{\prime}_{m},R). Hence for any xmB(xm,ω)x\notin\bigcup_{m}B(x_{m},\omega), we have Φ3(x)mB(pm,R)\Phi_{3}(x)\notin\bigcup_{m}B(p_{m}^{\prime},R), and again by the boundedness of dΦ31d\Phi_{3}^{-1} we have |g0ψ|2ρ>0\lvert\nabla_{g_{0}}\psi\rvert\geq 2\rho>0 for some uniform ρ\rho. As in (3.4), mB(xm,ω)Ω2β\bigcup_{m}B(x_{m},\omega)\subset\Omega_{2\beta}, the claim holds.

M\partial Mp1p_{1}^{\prime}x1x_{1}p2p_{2}^{\prime}x2x_{2}p3p_{3}^{\prime}x3x_{3}p4p_{4}^{\prime}x4x_{4}
Figure 2. Φ3\Phi_{3} moves the critical points of ψ2\psi_{2} into the sufficiently damped balls.

Remark 3.2.

The radial stretch Φ2\Phi_{2} is necessary here to pull the critical points sufficiently apart. Otherwise the balls B¯(pm,L+3ω)\bar{B}(p^{\prime}_{m},L+3\omega)’s in (3.16) may not be disjoint and the construction of the diffeomorphism Φ3\Phi_{3} fails in (3.18).

What makes this construction above interesting is that it is global on each cylindrical end, similar to the flavour of that on d\mathbb{R}^{d} 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 θk\theta_{k}’s to grow as rr\rightarrow\infty, for example, on conic ends, this construction stops working, technically because there is no ideal way of constructing a product-type ψ~2\tilde{\psi}_{2} 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 (Mk,g0)(M_{k},g_{0}), that is Mk=Mk×(1,)rM_{k}=\partial M_{k}\times(1,\infty)_{r} endowed with the metric g0=dr2+θ2k(r)hg_{0}=dr^{2}+\theta^{2}_{k}(r)h, where hh is a smooth metric on closed Mk\partial M_{k}, and θk(r)\theta_{k}(r) as described in (1.3). There exists some R1R\geq 1 and let MR=Mk×(R,)rM_{R}=\partial M_{k}\times(R,\infty)_{r}, and for some finite n1n\geq 1 there exist ψ1,,ψnCb(MR,)\psi_{1},\dots,\psi_{n}\in C^{\infty}_{b}(M_{R},\mathbb{R}) that each 0ψk30\leq\psi_{k}\leq 3, with a constant ρ>0\rho>0, such that for all kk, at each point xMRΩ2βx\in M_{R}\setminus\Omega_{2\beta} with |g0ψk(x)|<2ρ\lvert\nabla_{g_{0}}\psi_{k}(x)\rvert<2\rho, there is some ll depending on xx, such that ψl(x)1\psi_{l}(x)\geq 1 and

|g0ψl(x)|2ρ,ψl(x)ψk(x)+1/2.\lvert\nabla_{g_{0}}\psi_{l}(x)\rvert\geq 2\rho,\quad\psi_{l}(x)\geq\psi_{k}(x)+1/2. (3.19)
Proof.

In this lemma we write M=MkM=M_{k} and θ=θk\theta=\theta_{k} for we here only care what is happening on MkM_{k}.

EEMMΦΦk\Phi\circ\Phi_{k}
Figure 3. Placement of the critical points pulled back to MM.

1. We start by quasi-isometrically reducing the underlying geometry to an unbounded subset of d\mathbb{R}^{d}. As M\partial M is compact, it possesses a finite cover

Mk=1n(ϕ0k)1(B(0,1)),\partial M\subset\bigcup_{k=1}^{n}(\phi^{0}_{k})^{-1}(B(0,1)), (3.20)

such that each ϕ0k:M(ϕ0k)1(B(0,1))B(0,1)d1\phi^{0}_{k}:\partial M\supset(\phi^{0}_{k})^{-1}(B\left(0,1\right))\rightarrow B\left(0,1\right)\subset\mathbb{R}^{d-1} is a CbC_{b}^{\infty}-diffeomorphism. For convenience, denote for 0υ10\leq\upsilon\leq 1,

Συ=B(0,1υ),Συc=B(0,1)B(0,1υ),\Sigma_{\upsilon}=B(0,1-\upsilon),\quad\Sigma_{\upsilon}^{c}=B(0,1)\setminus B(0,1-\upsilon), (3.21)

where Συ\Sigma_{\upsilon} consists of points of distance more than υ\upsilon away from the unit sphere, Συc\Sigma_{\upsilon}^{c} is the υ\upsilon-neighbourhood of the unit sphere, and Σ0\Sigma_{0} is the unit open ball. Here, as the covers are open, one can fix a small ϵ>0\epsilon>0 such that for each kk, for each xx such that ϕ0k(x)Σc4ϵ\phi^{0}_{k}(x)\in\Sigma^{c}_{4\epsilon}, there is some ll such that x(ϕ0l)1(Σ4ϵ)x\in(\phi^{0}_{l})^{-1}\left(\Sigma_{4\epsilon}\right). Denote the model space by (D,gD)(D,g_{D}) where

D=Σ0×(1,)r,gD=θ2(r)dy2+dr2.D=\Sigma_{0}\times(1,\infty)_{r},\quad g_{D}=\theta^{2}(r)dy^{2}+dr^{2}. (3.22)

Construct diffeomorphisms Φk=ϕ0kIdr\Phi_{k}=\phi^{0}_{k}\otimes\operatorname{Id}_{r} and observe that {Φk1(D)}\{\Phi_{k}^{-1}(D)\} covers M×(1,)\partial M\times(1,\infty). Each map Φk\Phi_{k} is uniformly quasi-isometric with constants C1+,C1>0C_{1}^{+},C_{1}^{-}>0 such that

C1dM(x,x)dD(Φk(x),Φk(x))C1+dM(x,x),C_{1}^{-}d_{M}\left(x,x^{\prime}\right)\leq d_{D}\left(\Phi_{k}\left(x\right),\Phi_{k}\left(x^{\prime}\right)\right)\leq C_{1}^{+}d_{M}\left(x,x^{\prime}\right), (3.23)

for any x,xMx,x^{\prime}\in M. Consider Φ(y,r)=(θ(r)y,r)\Phi(y,r)=(\theta(r)y,r) for

Φ:(D,gD)(E,gd),E={(z,zd=r):r1,|z|θ(zd)}d.\Phi:\left(D,g_{D}\right)\rightarrow\left(E,g_{\mathbb{R}^{d}}\right),\quad E=\left\{(z^{\prime},z_{d}=r):r\geq 1,\lvert z^{\prime}\rvert\leq\theta(z_{d})\right\}\subset\mathbb{R}^{d}. (3.24)

This is a quasi-isometric CbC_{b}^{\infty}-diffeomorphism, that for any (y,r),(y,r)D(y,r),(y^{\prime},r^{\prime})\in D we have

C2dD((y,r),(y,r))dE(Φ((y,r)),Φ((y,r)))C2+dD((y,r),(y,r)).C_{2}^{-}d_{D}\left((y,r),(y^{\prime},r^{\prime})\right)\leq d_{E}\left(\Phi\left((y,r)\right),\Phi\left((y^{\prime},r^{\prime})\right)\right)\leq C_{2}^{+}d_{D}\left((y,r),(y^{\prime},r^{\prime})\right). (3.25)

To verify the CbC_{b}^{\infty} nature of Φ\Phi, it suffices to first pull back (D,gD)(D,g_{D}) to (E,gD)(E,g_{D}^{\prime}), and then verify that the Christoffel symbols on (E,gD)(E,g_{D}^{\prime}) are CbC_{b}^{\infty} on EE as a subset of d\mathbb{R}^{d}. We omit the trivial computation here. Note that each ΦΦk\Phi\circ\Phi_{k} is a CbC^{\infty}_{b}-diffeomorphism from ϕk1(D)\phi_{k}^{-1}(D) to EE quasi-isometric with constants C±=C±1C±2C^{\pm}=C^{\pm}_{1}C^{\pm}_{2}.

2. Now we construct on each EdE\subset\mathbb{R}^{d}, a weight function with vanishing gradients exactly inside the damping balls given in the Network Control Condition. On d\mathbb{R}^{d} we construct for k=1,,nk=1,\dots,n,

ψk0(z,zd)=cos(πzd2π(L+4ω)kC+2n(L+4ω)C+)j=1d1cos(πzj2(L+4ω)C+)+2,\psi_{k}^{0}(z^{\prime},z_{d})=\cos\left(\frac{\pi z_{d}-2\pi\left(L+4\omega\right)kC^{+}}{2n\left(L+4\omega\right)C^{+}}\right)\prod_{j=1}^{d-1}\cos{\left(\frac{\pi z_{j}}{2\left(L+4\omega\right)C^{+}}\right)}+2, (3.26)

whose critical points are

pm=(2C+(L+4ω)m,2C+(L+4ω)(nmd+k))p_{m}=\left(2C^{+}(L+4\omega)m^{\prime},2C^{+}(L+4\omega)(nm_{d}+k)\right) (3.27)

for all m=(m,md)dm=(m^{\prime},m_{d})\in\mathbb{Z}^{d}. See Figure 3. Note that any two such critical points are of distance at least 2C+(L+4ω)2C^{+}(L+4\omega), measured in d\mathbb{R}^{d}. Set R01R_{0}\geq 1 be the smallest constant such that for all rR0r\geq R_{0}, we have θ(r)>C1(L+4ω)/ϵ\theta(r)>C_{1}^{-}(L+4\omega)/\epsilon. This lower bound on the radius guarantees that any point in Φk1(Σϵ×(R0,))\Phi_{k}^{-1}(\Sigma_{\epsilon}\times(R_{0},\infty)) is of distance larger than L+4ωL+4\omega from the cross-sectional boundary Φk1(Σ0×(R0,))\Phi_{k}^{-1}(\partial\Sigma_{0}\times(R_{0},\infty)), measured in MM. By the Network Control Condition, for all critical points pmp_{m}’s that are inside Φ(Σϵ×(R0,))E\Phi(\Sigma_{\epsilon}\times(R_{0},\infty))\subset E there exists at least a xmx_{m} such that B(xm,ω)B((ΦΦk)1(pm),(L+2ω))B(x_{m},\omega)\subset B((\Phi\circ\Phi_{k})^{-1}(p_{m}),\left(L+2\omega\right)) and a2βa\geq 2\beta on B(xm,ω)B(x_{m},\omega). Here

ΦΦk(B(xm,ω))ΦΦk(B((ΦΦk)1(pm),(L+2ω)))B(pm,C+(L+2ω))\Phi\circ\Phi_{k}\left(B(x_{m},\omega)\right)\subset\Phi\circ\Phi_{k}\left(B((\Phi\circ\Phi_{k})^{-1}(p_{m}),\left(L+2\omega\right))\right)\\ \subset B(p_{m},C^{+}\left(L+2\omega\right)) (3.28)

are disjoint balls around pmp_{m}’s of some uniform radius. Hence via a process similar to the construction of the diffeomorphism Φ3\Phi_{3} in the proof of 3.1, we can find a CbC^{\infty}_{b}-diffeomorphism Φ~k\tilde{\Phi}_{k} on EE, equal to the identity on EmB(pm,C+(L+2ω))E\setminus\bigcup_{m}B(p_{m},C^{+}(L+2\omega)), such that Φ~k:Φk(xm)pm\tilde{\Phi}_{k}:\Phi_{k}(x_{m})\mapsto p_{m}. Set ψk1=Φ~kψk0\psi_{k}^{1}=\tilde{\Phi}_{k}^{*}\psi_{k}^{0}, whose critical points in Φ(Σϵ×(R0,))\Phi(\Sigma_{\epsilon}\times(R_{0},\infty)) are a subset of {Φk(xm)}\{\Phi_{k}(x_{m})\}. Set ψk2=Φψk1\psi_{k}^{2}=\Phi^{*}\psi_{k}^{1}. This is a function defined on Σ0×(R0,)D\Sigma_{0}\times(R_{0},\infty)\subset D. Note that 1ψ2k31\leq\psi^{2}_{k}\leq 3.

3. We now very carefully cut off the part of ψk2\psi_{k}^{2} within a small neighbourhood of Σ0×(R0,)\partial\Sigma_{0}\times(R_{0},\infty), and pull back and extend it to weight functions on MRM_{R} for some RR0R\geq R_{0}. Observe that away from Φk(nB(xm,ω))\Phi_{k}(\bigcup_{n}B(x_{m},\omega)) one has |gDψk2|C0\lvert\nabla_{g_{D}}\psi_{k}^{2}\rvert\geq C_{0} for some small C0C_{0}. Set RR0R\geq R_{0} to be that for all rRr\geq R, θ(r)36/C0\theta(r)\geq 36/C_{0}. Construct a cross-sectional cutoff χCc(Σ0)\chi\in C_{c}^{\infty}(\Sigma_{0}) such that χ(y)=0\chi(y)=0 on Σϵc\Sigma_{\epsilon}^{c}, greater than 1/121/12 on Σ2ϵ\Sigma_{2\epsilon}, less than 1/61/6 on Σ2ϵc\Sigma_{2\epsilon}^{c}, and identically 11 on Σ3ϵ\Sigma_{3\epsilon}. Moreover we ask |gDχ|C0/72\lvert\nabla_{g_{D}}\chi\rvert\leq C_{0}/72. See Figure 4. Note that we can find such a cutoff because RR is taken large enough to give the cross-section enough space to accommodate the tempered decay. Let the weight functions on Φk1(Σ0×(R,))MR\Phi_{k}^{-1}(\Sigma_{0}\times(R,\infty))\subset M_{R} be ψk=Φk(χ(y)ψ2k)\psi_{k}=\Phi_{k}^{*}(\chi(y)\psi^{2}_{k}). As χ(y)ψk2\chi(y)\psi_{k}^{2} is identically zero near Σ0×(R,)\partial\Sigma_{0}\times(R,\infty), we extend ψk\psi_{k} to all of MRM_{R} by 0. Note that in general 0ψk30\leq\psi_{k}\leq 3, and specifically on Φk1(Σ4ϵ×(R,))\Phi_{k}^{-1}(\Sigma_{4\epsilon}\times(R,\infty)) we have 1ψk31\leq\psi_{k}\leq 3 .

Σ0\partial\Sigma_{0}ϵ\epsilon2ϵ2\epsilon3ϵ3\epsilon4ϵ4\epsilon1/121/61
Figure 4. Behaviour of χ(y)\chi(y) near Σ0\partial\Sigma_{0}.

4. We claim that ψk\psi_{k}’s meet our requirement listed in the statement. There is a lower bound for the pushforward map dΦkC1\|d\Phi_{k}\|\geq C_{1}. Hence we have

|g0ψk|C1|gDχψ2k|.\lvert\nabla_{g_{0}}\psi_{k}\rvert\geq C_{1}\lvert\nabla_{g_{D}}\chi\psi^{2}_{k}\rvert. (3.29)

Fix kk and some point xMRx\in M_{R} while xnB(xm,ω)x\notin\bigcup_{n}B(x_{m},\omega). Note this means Φk(x)Σ0×(R,)\Phi_{k}(x)\in\Sigma_{0}\times(R,\infty) and Φk(x)Φk(nB(xm,ω))\Phi_{k}(x)\notin\Phi_{k}(\bigcup_{n}B(x_{m},\omega)). Now set 2ρ=C0C1/242\rho=C_{0}C_{1}/24. There are three circumstances depending where Φk(x)\Phi_{k}(x) is.

(a) If Φk(x)Σ3ϵ×(R,)\Phi_{k}(x)\in\Sigma_{3\epsilon}\times(R,\infty), the cross-sectional cutoff χ\chi is identically 11. We have

|gDχψk2|=|gDψk2|C0,\lvert\nabla_{g_{D}}\chi\psi_{k}^{2}\rvert=\lvert\nabla_{g_{D}}\psi_{k}^{2}\rvert\geq C_{0}, (3.30)

and therefore |g0ψk|C0C1=48ρ2ρ\lvert\nabla_{g_{0}}\psi_{k}\rvert\geq C_{0}C_{1}=48\rho\geq 2\rho.

(b) If Φk(x)(Σ2ϵΣc3ϵ)×(R,)\Phi_{k}(x)\in(\Sigma_{2\epsilon}\bigcap\Sigma^{c}_{3\epsilon})\times(R,\infty), the cross-sectional cutoff χ>1/12\chi>1/12 is sufficiently large. We have

|gDχψ2k||χ||gDψ2k||ψ2k||gDχ|C0/123C0/72=C0/24.\lvert\nabla_{g_{D}}\chi\psi^{2}_{k}\rvert\geq\lvert\chi\rvert\lvert\nabla_{g_{D}}\psi^{2}_{k}\rvert-\lvert\psi^{2}_{k}\rvert\lvert\nabla_{g_{D}}\chi\rvert\geq C_{0}/12-3C_{0}/72=C_{0}/24. (3.31)

Here we used the fact that ψk\psi_{k} is bounded from above by 3. Therefore |g0ψk|C0C1/24=2ρ\lvert\nabla_{g_{0}}\psi_{k}\rvert\geq C_{0}C_{1}/24=2\rho.

(c) If Φk(x)Σ2ϵc×(R,)\Phi_{k}(x)\in\Sigma_{2\epsilon}^{c}\times(R,\infty), there is some lnl\leq n such that Φl(x)Σ4ϵ×(R,)\Phi_{l}(x)\in\Sigma_{4\epsilon}\times(R,\infty). From the circumstance (a), we know that |g0ψl(x)|2ρ\lvert\nabla_{g_{0}}\psi_{l}(x)\rvert\geq 2\rho, and

ψk(x)=χψ2k(ϕk(x))1/2=11/2ψl(x)1/2.\psi_{k}(x)=\chi\psi^{2}_{k}(\phi_{k}(x))\leq 1/2=1-1/2\leq\psi_{l}(x)-1/2. (3.32)

Here we used the fact that χ1/6,ψk23\chi\leq 1/6,\psi_{k}^{2}\leq 3 and ψl1\psi_{l}\geq 1.

The claim has been concluded as above. ∎

Remark 3.4.
  1. (1)

    We note that the argument is sharp for conic ends, where θ(r)=r\theta(r)=r. If θ(r)\theta^{\prime}(r) is not uniformly bounded, then Φ\Phi loses the quasi-isometric nature, and the argument needs further modifications.

  2. (2)

    This argument relies, twice when setting up R0R_{0} and RR, on the fact that the cross-sectional space is expanding as rr\rightarrow\infty. Large R0R_{0} makes sure that the Φk1(Σϵ×(R0,))\Phi_{k}^{-1}(\Sigma_{\epsilon}\times(R_{0},\infty)) is sufficiently apart from Φk1(Σ0×(R0,))\Phi_{k}^{-1}(\partial\Sigma_{0}\times(R_{0},\infty)), so the critical points inside Φk1(Σϵ×(R0,))\Phi_{k}^{-1}(\Sigma_{\epsilon}\times(R_{0},\infty)) will not be pulled to some xmx_{m} out of the charted region Φk1(Σ0×(R0,))\Phi_{k}^{-1}(\Sigma_{0}\times(R_{0},\infty)). 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 (M,g0)(M,g_{0}) 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 (M,g)(M,g).

Proposition 3.5 (Construction of Carleman weights).

On the prespecified (M,g)(M,g), there are Carleman weights ψ0,,ψnCb(M)\psi_{0},\dots,\psi_{n}\in C_{b}^{\infty}(M) compatible with control from (Ωβ,Ω2β)(\Omega_{\beta},\Omega_{2\beta}), in the sense of (2.1).

Proof.

1. We start by reviewing what we have learnt from the previous lemmata. Denote {xk(Mk×(a,b)r)}\{x\in\bigcup_{k}(\partial M_{k}\times(a,b)_{r})\} by (a,b)(a,b), its closure by [a,b][a,b], and M0(1,b)M_{0}\cup(1,b) by {r<b}\{r<b\} as a matter of convenience. Lemma 3.1 and Lemma 3.3 state that we have an uniform R1R\geq 1 with a finite family of weights ψ1,,ψn\psi_{1}^{\prime},\dots,\psi_{n}^{\prime} on the ends (R,)(R,\infty), where 0ψl30\leq\psi_{l}^{\prime}\leq 3 on the end Mkl×(R,)\partial M_{k_{l}}\times(R,\infty) where it is defined, and 1ψl31\leq\psi_{l}^{\prime}\leq 3 specifically on some UlMkl×(R,)U_{l}\subset\partial M_{k_{l}}\times(R,\infty), with (R,)Ω2βcl=1nUl(R,\infty)\cap\Omega_{2\beta}^{c}\subset\bigcup_{l=1}^{n}U_{l}. Moreover there exists ρ1>0\rho_{1}>0, such that for each ll, at xUlΩ2βx\in U_{l}\setminus\Omega_{2\beta}, one has |g0ψl|ρ1\lvert\nabla_{g_{0}}\psi_{l}^{\prime}\rvert\geq\rho_{1}, and if for some kk we have |g0ψk|<ρ1\lvert\nabla_{g_{0}}\psi_{k}^{\prime}\rvert<\rho_{1} then ψlψk+1/2\psi_{l}\geq\psi_{k}+1/2.

2. On (M,g0)(M,g_{0}), we start by constructing a weight on the central compactum, to which the ends on which we have the weights are attached. Set I=432/ρ1I=432/\rho_{1}. On M0={rR+7I}M_{0}^{\prime}=\{r\leq R+7I\} compact, there exists a Morse function ψ01\psi_{0}^{1} with finitely many non-degenerate critical points, none of which resides on the boundary {r=R+7I}\{r=R+7I\}. Note that this can be achieved by finding a Morse function on a small closed neighbourhood of M0M_{0}^{\prime}, for example {rR+8I}\{r\leq R+8I\}, and find a diffeomorphism to move all critical points not on the new boundary {r=R+8I}\{r=R+8I\} into {r<R+7I}\{r<R+7I\} and then restrict the new function to {rR+7I}\{r\leq R+7I\}. Apply a diffeomorphism on M0M_{0}^{\prime}, to get ψ02\psi_{0}^{2} where all critical points of ψ10\psi^{1}_{0} are moved inside some B(x0,ω)B(x_{0},\omega) given by the Network Control Condition. Note that we can assume without loss of generality that there is a such B(x0,ω)B(x_{0},\omega) inside M0M_{0}^{\prime}, by increasing RR if needed. Now construct

ψ0=13+13(maxψ02minψ02)1(ψ02minψ02).\psi_{0}^{\prime}=\frac{1}{3}+\frac{1}{3}\left(\max{\psi_{0}^{2}}-\min{\psi_{0}^{2}}\right)^{-1}\left(\psi_{0}^{2}-\min{\psi_{0}^{2}}\right). (3.33)

Note that ψ0[1/3,2/3]\psi_{0}^{\prime}\in[1/3,2/3] on M0M_{0}^{\prime}, and |g0ψ0|ρ0\lvert\nabla_{g_{0}}\psi_{0}^{\prime}\rvert\geq\rho_{0} for some positive ρ0\rho_{0} away from B(x0,ω)Ω2βB(x_{0},\omega)\subset\Omega_{2\beta}. Now set

2ρ=min{ρ0,ρ1/72}.2\rho=\min\{\rho_{0},\rho_{1}/72\}. (3.34)

We have |g0ψ0|2ρ\lvert\nabla_{g_{0}}\psi_{0}^{\prime}\rvert\geq 2\rho on M0Ω2βM_{0}^{\prime}\setminus\Omega_{2\beta}, and ρ1144ρ\rho_{1}\geq 144\rho. These are weight functions on (M,g0)(M,g_{0}).

3. We now trim the parts of ψ0,,ψn\psi_{0}^{\prime},\dots,\psi_{n}^{\prime} inside [R,R+6I][R,R+6I], where the supports of those weights could intersect, and extend them to the whole (M,g0)(M,g_{0}) in a compatible manner. Construct two radial cutoff functions χ0\chi_{0} and χ1\chi_{1} in Cb(M)C^{\infty}_{b}(M). Let χ0(r)\chi_{0}(r) be non-increasing, 11 on {rR+4I}\{r\leq R+4I\}, and 0 on [R+5I,)[R+5I,\infty). Let χ1(r)\chi_{1}(r) be non-decreasing, 0 on {rR+I}\{r\leq R+I\}, and χ11/36\chi_{1}\geq 1/36 on [R+2I,)[R+2I,\infty), and χ11/18\chi_{1}\leq 1/18 on [R,R+2I][R,R+2I], and constant 11 on [R+3I,)[R+3I,\infty). Meanwhile we ask |rχ1|ρ1/216\lvert\partial_{r}\chi_{1}\rvert\leq\rho_{1}/216 on [R,)[R,\infty), which makes sense as I=432/ρ1I=432/\rho_{1} has been taken large enough. See Figure LABEL:F2. Now set

ψ0=χ0ψ0,ψk=χ1ψk,\psi_{0}=\chi_{0}\psi_{0}^{\prime},\quad\psi_{k}=\chi_{1}\psi_{k}^{\prime}, (3.35)

extended to the whole manifold MM by 0. Note ψ0,,ψn\psi_{0},\dots,\psi_{n} are in Cb(M)C_{b}^{\infty}(M).

RRR+IR+IR+2IR+2IR+3IR+3IR+4IR+4IR+5IR+5IR+6IR+6I1/361/181
RRR+IR+IR+2IR+2IR+3IR+3IR+4IR+4IR+5IR+5IR+6IR+6I