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

Positive mass and Dirac operators on weighted manifolds and smooth metric measure spaces

Michael B. Law Isaac M. Lopez  and  Daniel Santiago MIT, Department of Mathematics, 77 Massachusetts Avenue, Cambridge, MA 02139, USA. mikelaw@mit.edu, imlopez@mit.edu, dsantiag@mit.edu
Abstract.

We establish a weighted positive mass theorem which unifies and generalizes results of Baldauf–Ozuch and Chu–Zhu. Our result is in fact equivalent to the usual positive mass theorem, and can be regarded as a positive mass theorem for smooth metric measure spaces. We also study Dirac operators on certain warped product manifolds associated to smooth metric measure spaces. Applications of this include, among others, an alternative proof for a special case of our positive mass theorem, eigenvalue bounds for the Dirac operator on closed spin manifolds, and a new way to understand the weighted Dirac operator using warped products.

1. Introduction

One of the crowning achievements in contemporary geometry is the positive mass theorem in general relativity. We recall the Riemannian version of this theorem, referring to §2.1 for the definitions of asymptotically Euclidean (AE) manifolds and ADM mass used in this paper.

Theorem 1.1 ([33, 34, 36]).

Let (Mn,g)(M^{n},g), n3n\geq 3 be an AE manifold of order τ>n22\tau>\frac{n-2}{2}, and assume that 3n73\leq n\leq 7 or MM is spin. If (Mn,g)(M^{n},g) has nonnegative scalar curvature R0R\geq 0, then it has nonnegative ADM mass 𝔪(g)0\mathfrak{m}(g)\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}).

The positive mass theorem was proved by Schoen and Yau in the 3n73\leq n\leq 7 case and by Witten in the spin case.111Mathematical accounts of Witten’s proof can be found in [25], or in [30] for the spacetime case. Among the many related developments that have since arisen, the most relevant to this paper are the generalizations of the positive mass theorem to weighted manifolds. These were recently established by Baldauf–Ozuch [6] in the spin case and Chu–Zhu [12] in the 3n73\leq n\leq 7 case.

A weighted manifold (Mn,g,f)(M^{n},g,f) is a Riemannian manifold (Mn,g)(M^{n},g) with a weight f𝒞(M)f\in\mathcal{C}^{\infty}(M) which defines the measure efdVolge^{-f}d\mathrm{Vol}_{g}. They were first studied by Lichnerowicz [26, 27] and appear in many parts of mathematics, such as in Ricci flow thanks to the work of Perelman [31]. Perelman characterized Ricci flow as the gradient flow of the functional \mathcal{F} defined on weighted manifolds by

(1.1) (M,g,f):=MRfef𝑑Volg,\mathcal{F}(M,g,f):=\int_{M}R_{f}e^{-f}d\mathrm{Vol}_{g},

where Rf=R+2Δf|f|2R_{f}=R+2\Delta f-|\nabla f|^{2} is the weighted scalar curvature (or PP-scalar curvature). Besides, weighted manifolds find applications in physics through the Brans–Dicke theory of scalar-tensor gravitation [37, 21] as well as theories involving Kaluza–Klein compactifications [14].

Various facts about manifolds with positive (resp. nonnegative) scalar curvature are known to generalize to weighted manifolds with positive (resp. nonnegative) weighted scalar curvature. Results of this flavor include those of e.g. [18, 1, 15], as well as the weighted positive mass theorem described next. In [6], Baldauf and Ozuch define the weighted mass 𝔪f(g)\mathfrak{m}_{f}(g) of an AE weighted manifold (Mn,g,f)(M^{n},g,f) by

(1.2) 𝔪f(g):=𝔪(g)+2limρSρf,𝐧ef𝑑A,\mathfrak{m}_{f}(g):=\mathfrak{m}(g)+2\lim_{\rho\rightarrow\infty}\int_{S_{\rho}}\langle\nabla f,\mathbf{n}\rangle e^{-f}dA,

where SρS_{\rho} is a coordinate sphere in the end of MM, 𝐧\mathbf{n} is the Euclidean outward unit normal, and dAdA is the Euclidean area element. The weighted positive mass theorem reads as follows. (The weighted Hölder spaces 𝒞βk,α(M)\mathcal{C}^{k,\alpha}_{\beta}(M) are defined in Definition 2.3.)

Theorem 1.2 ([6, 12]).

Let (Mn,g,f)(M^{n},g,f), n3n\geq 3 be an AE weighted manifold of order τ>n22\tau>\frac{n-2}{2}, and assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g).

  1. (a)

    Suppose MM is spin and Rf0R_{f}\geq 0. Then 𝔪f(g)0\mathfrak{m}_{f}(g)\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}) and n(Δff)ef𝑑x=0\int_{\mathbb{R}^{n}}(\Delta_{f}f)e^{-f}\,dx=0.

  2. (b)

    Suppose 3n73\leq n\leq 7 and Rf0R_{f}\geq 0, with slightly more decay on ff and gg (omitting details). Then 𝔪f(g)0\mathfrak{m}_{f}(g)\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}) and f0f\equiv 0.

Baldauf and Ozuch prove (a) by adapting Witten’s proof to the weighted setting, while Chu and Zhu prove (b) by adapting Schoen and Yau’s argument. We elaborate on the former for the sake of later discussion. Recall that Witten’s proof hinges on finding a spinor ϕ\phi on MM such that

  1. (i)

    Dϕ=0D\phi=0, where DD is the Dirac operator.

  2. (ii)

    ϕ\phi asymptotes to a constant unit norm spinor at infinity.

A spinor satisfying (i) and (ii) is called a Witten spinor. These properties are used to prove Witten’s formula for the mass in terms of ϕ\phi and the scalar curvature RR:

(1.3) 𝔪(g)=4M(|ϕ|2+14R|ϕ|2)𝑑Volg.\mathfrak{m}(g)=4\int_{M}\left(|\nabla\phi|^{2}+\frac{1}{4}R|\phi|^{2}\right)d\mathrm{Vol}_{g}.

Thus 𝔪(g)0\mathfrak{m}(g)\geq 0 if R0R\geq 0. Baldauf and Ozuch proceed similarly to prove Theorem 1.2(a); they find a weighted Witten spinor ϕ\phi with properties analogous to (i) and (ii):

  1. (i’)

    Dfϕ=0D_{f}\phi=0, where DfD_{f} is the weighted Dirac operator

    (1.4) Df=D12f.D_{f}=D-\frac{1}{2}\nabla f\cdot.
  2. (ii’)

    ϕ\phi asymptotes to a constant unit norm spinor at infinity.

These properties lead to a formula for the weighted mass analogous to (1.3):

(1.5) 𝔪f(g)=4M(|ϕ|2+14Rf|ϕ|2)ef𝑑Volg,\mathfrak{m}_{f}(g)=4\int_{M}\left(|\nabla\phi|^{2}+\frac{1}{4}R_{f}|\phi|^{2}\right)e^{-f}d\mathrm{Vol}_{g},

from which Theorem 1.2(a) follows.

We now introduce the main results of this paper.

1.1. Equivalence between weighted and unweighted positive mass theorems

The unweighted positive mass theorem (Theorem 1.1) is the f=0f=0 case of the weighted positive mass theorem (Theorem 1.2). Our first result says that these theorems are actually equivalent. We also sharpen the conclusions of Theorem 1.2:

Theorem 1.3.

Let (Mn,g,f)(M^{n},g,f), n3n\geq 3 be an AE weighted manifold of order τ>n22\tau>\frac{n-2}{2}, and assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g). Also suppose 3n73\leq n\leq 7 or MM is spin.

  1. (a)

    If Rf0R_{f}\geq 0, then 𝔪f(g)0\mathfrak{m}_{f}(g)\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}) and f0f\equiv 0.

  2. (b)

    The result of part (a) is equivalent to the unweighted positive mass theorem (Theorem 1.1).

  3. (c)

    If Rf1n1|f|2R_{f}\geq-\frac{1}{n-1}|\nabla f|^{2}, then we still have 𝔪f(g)0\mathfrak{m}_{f}(g)\geq 0.

Theorem 1.3(a) improves on Theorem 1.2 due to the sharper rigidity in the spin case. Theorem 1.3(b) follows from the proof of part (a), which uses a suitable conformal change of metric to reduce to the unweighted positive mass theorem. Theorem 1.3(c) follows directly from the associated computations, and further strengthens the result of part (a) by weakening the lower bound on RfR_{f} that guarantees nonnegativity of the weighted mass.

1.2. A positive mass theorem for smooth metric measure spaces

Generalizing beyond weighted manifolds, one arrives at smooth metric measure spaces (SMMSs). The following definition is from [11], which introduces SMMSs more thoroughly and unifies the perspectives of Bakry–Émery, Chang–Gursky–Yang, and Perelman on the subject.

Definition 1.4.

A smooth metric measure space (SMMS) is a 4-tuple =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) where (Mn,g)(M^{n},g) is a Riemannian manifold, efdVolge^{-f}d\mathrm{Vol}_{g} is a measure defined by a weight f𝒞(M)f\in\mathcal{C}^{\infty}(M), and mm\in\mathbb{R}.

The analogs of Ricci and scalar curvatures on an SMMS with m0m\neq 0 are the mm-Bakry–Émery Ricci and scalar curvatures, respectively

(1.6) Ricfm\displaystyle\mathrm{Ric}^{m}_{f} :=Ric+Hessf1mdfdf,\displaystyle:=\mathrm{Ric}+\mathrm{Hess}_{f}-\frac{1}{m}df\otimes df,
(1.7) Rfm\displaystyle R^{m}_{f} :=Rf1m|f|2=R+2Δfm+1m|f|2.\displaystyle:=R_{f}-\frac{1}{m}|\nabla f|^{2}=R+2\Delta f-\frac{m+1}{m}|\nabla f|^{2}.

As mm\to\infty, we have RfmRfR^{m}_{f}\to R_{f} and RicfmRicf\mathrm{Ric}^{m}_{f}\to\mathrm{Ric}_{f}, where Ricf=Ric+Hessf\mathrm{Ric}_{f}=\mathrm{Ric}+\mathrm{Hess}_{f} is commonly called the Bakry–Émery Ricci curvature. Because RfR_{f} and Ricf\mathrm{Ric}_{f} typically appear in the context of weighted manifolds (M,g,f)(M,g,f), e.g. in Ricci flow, weighted manifolds can be viewed as SMMSs with m=m=\infty. Other values of mm bear significance in geometry and physics (see e.g. [11, 38] and [39, pp.1081–82]). Of special importance to us are SMMSs with mm\in\mathbb{N}, in which case Ricfm\mathrm{Ric}^{m}_{f} and RfmR^{m}_{f} arise from certain warped products over (M,g)(M,g). Specifically, if (Fm,h)(F^{m},h) is an mm-dimensional scalar-flat manifold, then the warped product (Mn×Fm,g¯=ge2fmh)(M^{n}\times F^{m},\bar{g}=g\oplus e^{-\frac{2f}{m}}h) has scalar curvature RfmR_{f}^{m}, and its Ricci tensor satisfies Ricg¯(X,Y)=Ricfm(X,Y)\mathrm{Ric}_{\overline{g}}(X,Y)=\mathrm{Ric}_{f}^{m}(X,Y) for all vector fields X,YTMX,Y\in TM [8, Proposition 9.106].

Definition 1.5.

An SMMS =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) is called asymptotically Euclidean (AE) if (Mn,g)(M^{n},g) is AE. If \mathcal{M} is an AE SMMS, then its mass is defined to simply be the weighted mass of (Mn,g,f)(M^{n},g,f):

(1.8) 𝔪():=𝔪f(g).\mathfrak{m}(\mathcal{M}):=\mathfrak{m}_{f}(g).

The mass is curiously independent of mm, but is motivated as follows. If mm\in\mathbb{N}, then there is a close connection between the SMMS \mathcal{M} and the warped product (Mn×Fm,g¯=ge2fmh)(M^{n}\times F^{m},\bar{g}=g\oplus e^{-\frac{2f}{m}}h) as discussed above. Normalizing so that (Fm,h)(F^{m},h) has unit volume, it is natural to define the mass of \mathcal{M} similarly to the mass of an AE manifold:

(1.9) 𝔪():=limρSρM×F(ig¯ijjg¯aa)𝐧j𝑑A𝑑Volh.\mathfrak{m}(\mathcal{M}):=\lim_{\rho\to\infty}\int_{S^{M}_{\rho}\times F}(\partial_{i}\bar{g}_{ij}-\partial_{j}\bar{g}_{aa})\mathbf{n}_{j}\,dA\,d\mathrm{Vol}_{h}.

(See Definition 2.6 for a precise definition.) We will show that this coincides with the weighted mass 𝔪f(g)\mathfrak{m}_{f}(g). This leads us to define the mass of an AE SMMS, with mm not necessarily in \mathbb{N}, as its weighted mass.

As RfmR^{m}_{f} (1.7) is the analog of scalar curvature on an SMMS, a positive mass theorem for SMMSs should assert that an AE SMMS with Rfm0R^{m}_{f}\geq 0 has 𝔪()0\mathfrak{m}(\mathcal{M})\geq 0. Using Theorem 1.3, we will see that such a theorem does indeed hold for mm outside the interval (1n,0](1-n,0]. This is our second main result.

Theorem 1.6.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be an AE SMMS of order τ>n22\tau>\frac{n-2}{2} with m(1n,0]m\in\mathbb{R}\setminus(1-n,0], and suppose 3n73\leq n\leq 7 or MM is spin. Also assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g). If Rfm0R^{m}_{f}\geq 0, then 𝔪()0\mathfrak{m}(\mathcal{M})\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}) and f0f\equiv 0.

Remark.

In [13], Dai proved a positive mass theorem for spin manifolds asymptotic to (nBR(0)¯)×X(\mathbb{R}^{n}\setminus\overline{B_{R}(0)})\times X, where XX is a Calabi–Yau (hence scalar-flat) manifold. Dai defines a mass for such manifolds, which coincides with (1.9) up to a factor if the manifold is globally the product of an AE manifold and XX. Therefore, Theorem 1.6 implies Dai’s positive mass theorem in this special case. The advantange of this approach is that it avoids the Mazzeo–Melrose fibered boundary calculus that Dai used to prove the general case of his result.

1.3. Warped product Dirac operators and applications

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be a SMMS with mm\in\mathbb{N}, and suppose MM is spin. Let (Fm,h)(F^{m},h) be a spin manifold and form the warped product

(1.10) (Mn×Fm,g¯=ge2fmh).(M^{n}\times F^{m},\overline{g}=g\oplus e^{-\frac{2f}{m}}h).

Using [32, §3.2], we will identify the spinor bundle Σ¯(M×F)\bar{\Sigma}(M\times F) of the warped product in terms of the spinor bundles ΣM\Sigma M and ΣF\Sigma F, and describe the Dirac operator on Σ¯(M×F)\bar{\Sigma}(M\times F) in terms of operators on ΣM\Sigma M and ΣF\Sigma F. For illustration purposes, assume that MM is even-dimensional. We will see that Σ¯(M×F)ΣMΣF\bar{\Sigma}(M\times F)\cong\Sigma M\otimes\Sigma F as vector bundles, and that the Dirac operator on Σ¯(M×F)\bar{\Sigma}(M\times F) acts in the following way. The next theorem is a special case of Theorem 3.5, our third main result.

Theorem 1.7.

Let MM be even-dimensional, ϕΓ(ΣM)\phi\in\Gamma(\Sigma M) be a spinor on MM and νΓ(ΣF)\nu\in\Gamma(\Sigma F) be a parallel spinor, i.e. ν=0\nabla\nu=0. Then the Dirac operator D¯\bar{D} on Σ¯(M×F)ΣMΣF\bar{\Sigma}(M\times F)\cong\Sigma M\otimes\Sigma F satisfies

(1.11) D¯(ϕν)=(Dfϕ)ν,\bar{D}(\phi\otimes\nu)=(D_{f}\phi)\otimes\nu,

where DfD_{f} is the weighted Dirac operator on ΣM\Sigma M, defined in (1.4).

Thus, the weighted Dirac operator DfD_{f} arises naturally from the SMMS \mathcal{M} associated to MM. This stands in contrast to the motivations for DfD_{f} provided by Perelman [31] and Baldauf–Ozuch [6] (where DfD_{f} is an ad hoc operator allowing identities like the Lichnerowicz formula to be generalized to weighted manifolds), and by Branding and Habib [10] (who exhibit DfD_{f} as the Euler–Lagrange operator of a spinorial energy involving the weighted measure efdVolge^{-f}d\mathrm{Vol}_{g}).

Intriguingly, the fiber dimension m=dimFm=\dim F does not appear in Theorem 1.7. This may have interesting geometric consequences beyond this paper, and possibly consequences in physics too. If FF has special holonomy except quaternionic Kähler, then it has nontrivial parallel spinors [35], so Theorem 1.7 can be meaningfully applied to the warped product (M×F,g¯)(M\times F,\bar{g}). Such products, and more generally geometries with dimensions ‘hidden’ in special holonomy fibers, are frequently encountered in string theories, so the fiber-dimension independence in Theorem 1.7 might prove significant in those contexts. We remark that Dirac operators on product manifolds have been investigated in the setting of collapsing fibrations [2, 28, 32] which is the essence of Kaluza–Klein dimensional reduction. However, to our knowledge, existing work does not examine the dependence of the results (or lack thereof) on the fiber dimension.

This paper gives several applications of Theorem 1.7 and its surrounding computations. Firstly, we will use Theorem 1.7 to turn facts from unweighted spin geometry on the warped product into facts in weighted spin geometry on MM. This allows us to systematically reprove some known identities in weighted spin geometry, one being the weighted Witten formula (1.5) which in turn gives a second proof of a special case of Theorem 1.6. Other applications pertain to the spectrum of the Dirac operator on closed manifolds, such as using Theorem 1.7 to give eigenvalue bounds in terms of RfmR_{f}^{m}. Moreover, we will leverage a relationship between the weighted Dirac operator DfD_{f} and the conformal metric used to prove Theorem 1.3 to generalize the classical fact that a closed spin manifold with positive scalar curvature admits no nontrivial harmonic spinors:

Corollary 1.8.

Let MM be a closed spin manifold. If for some m[1n,0]m\in\mathbb{R}\setminus[1-n,0] we have Rfm0R^{m}_{f}\geq 0 and Rfm>0R^{m}_{f}>0 at some point, then MM admits no nontrivial harmonic spinors.

For each m{0}m\in\mathbb{R}\setminus\{0\}, there is a unique μm\mu_{m}\in\mathbb{R} such that Rfm=μmR^{m}_{f}=\mu_{m} has a solution f𝒞(M)f\in\mathcal{C}^{\infty}(M). While we are unable to combine this with Corollary 1.8 to get new obstructions to harmonic spinors, we find that the μm\mu_{m} yield a family of inequalities λ1(D)2n4(n1)μm\lambda_{1}(D)^{2}\geq\frac{n}{4(n-1)}\mu_{m} for the lowest eigenvalue of the Dirac operator, interpolating between the (stronger) Friedrich and Hijazi inequalities [19, 22] as mm varies.

Organization

In Section 2, after stating definitions and conventions, we prove our positive mass theorems, Theorem 1.3 and Theorem 1.6. We also motivate our definition of mass for SMMSs. In Section 3, we study the spin geometry of warped products, eventually relating the connection and Dirac operator on (M×F,g¯)(M\times F,\bar{g}) to the corresponding objects on MM and FF in Theorem 3.5. In Section 4 we discuss applications of Theorem 3.5 (or its special case Theorem 1.7) and other computations, as outlined above.

Acknowledgments

The authors thank Tristan Ozuch for suggesting the project, and for continuously giving invaluable insights and advice. M.L. was supported in part by a Croucher Scholarship. I.M.L. and D.S. were supported in part by the MIT Department of Mathematics through its Summer Program in Undergraduate Research (SPUR).

2. Positive mass theorems

We begin in §2.1 by stating our conventions for AE manifolds and ADM mass. We then prove our positive mass theorems, Theorem 1.3 and Theorem 1.6, in §2.2 and §2.4 respectively. In §2.3, we motivate the weighted mass 𝔪f(g)\mathfrak{m}_{f}(g) as a reasonable notion of mass for SMMSs.

2.1. Asymptotically Euclidean manifolds and mass

We adhere to the following conventions in this paper. A detailed exposition to the concepts below may be found in [24].

Definition 2.1.

A complete Riemannian manifold (Mn,g)(M^{n},g) of dimension n3n\geq 3 is said to be asymptotically Euclidean (AE) of order τ>n22\tau>\frac{n-2}{2} if

  1. (a)

    There is a decomposition M=McpctMM=M_{\mathrm{cpct}}\cup M_{\infty}, where McpctM_{\mathrm{cpct}} is compact and MM_{\infty} is diffeomorphic to the complement of a closed ball in n\mathbb{R}^{n}.

  2. (b)

    In the induced asymptotic coordinates xx for MM_{\infty}, we have for ρ=|x|\rho=|x| the falloff conditions

    (2.1) gij=δij+𝒪(ρτ),kgij=𝒪(ρτ1),klgij=𝒪(ρτ2).g_{ij}=\delta_{ij}+\mathcal{O}(\rho^{-\tau}),\quad\partial_{k}g_{ij}=\mathcal{O}(\rho^{-\tau-1}),\quad\partial_{k}\partial_{l}g_{ij}=\mathcal{O}(\rho^{-\tau-2}).
  3. (c)

    The scalar curvature R=RgR=R_{g} belongs to L1(M,g)L^{1}(M,g).

While this definition excludes the possibility of having multiple ends, the results of this paper extend to that case upon applying now-standard modifications.

Definition 2.2 ([3, 4, 5]).

The (ADM) mass of an AE manifold (M,g)(M,g) is

(2.2) 𝔪(g):=limρSρi,j=1n(igijjgii)𝐧jdA,\mathfrak{m}(g):=\lim_{\rho\rightarrow\infty}\int_{S_{\rho}}\sum_{i,j=1}^{n}(\partial_{i}g_{ij}-\partial_{j}g_{ii})\mathbf{n}_{j}\,dA,

where SρS_{\rho} is the coordinate sphere of radius ρ\rho in asymptotic coordinates MnBR(0)¯M_{\infty}\cong\mathbb{R}^{n}\setminus\overline{B_{R}(0)}, 𝐧j\mathbf{n}_{j} is the jj-th component of the outward Euclidean unit normal to SρS_{\rho}, and dAdA is the Euclidean area element.

According to [7] (see also [25, §9]), the integral (2.2) is finite on an AE manifold and does not depend on the choice of asymptotic coordinates; thus the mass is an invariant of gg. The mass is defined with different normalizing constants in various references; we have chosen to follow the convention of [6].

Definition 2.3.

Let (Mn,g)(M^{n},g) be an AE manifold with asymptotic coordinates xx on MM_{\infty}, as in Definition 2.1. For 0<α<10<\alpha<1, k0k\in\mathbb{N}_{0} and β\beta\in\mathbb{R}, the weighted Hölder space 𝒞βk,α\mathcal{C}^{k,\alpha}_{\beta} is the space of 𝒞k\mathcal{C}^{k} functions u:Mu:M\to\mathbb{R} for which the norm

(2.3) u𝒞βk,α(M):=0ik(supxM|iu(x)||x|βi)+supxM[ku]𝒞α(B|x|/2(x))|x|β(k+α)\lVert u\rVert_{\mathcal{C}^{k,\alpha}_{\beta}(M)}:=\sum_{0\leq i\leq k}\left(\sup_{x\in M_{\infty}}\frac{|\nabla^{i}u(x)|}{|x|^{\beta-i}}\right)+\sup_{x\in M_{\infty}}\frac{[\nabla^{k}u]_{\mathcal{C}^{\alpha}(B_{|x|/2}(x))}}{|x|^{\beta-(k+\alpha)}}

is finite, where B|x|/2(x)B_{|x|/2}(x) is the metric ball of radius |x|2\frac{|x|}{2} centered at xx and

(2.4) [ku]𝒞α(B|x|/2(x))\displaystyle[\nabla^{k}u]_{\mathcal{C}^{\alpha}(B_{|x|/2}(x))} :=supy,zB|x|2(x)|ku(y)ku(z)||yz|α.\displaystyle:=\sup_{y,z\in B_{\frac{|x|}{2}}(x)}\frac{|\nabla^{k}u(y)-\nabla^{k}u(z)|}{|y-z|^{\alpha}}.

If EE is a smooth vector bundle over MM equipped with a bundle metric and connection, then the spaces of sections 𝒞βk,α(E)\mathcal{C}^{k,\alpha}_{\beta}(E) are defined analogously.

Note that if u𝒞βk,α(M)u\in\mathcal{C}^{k,\alpha}_{\beta}(M), then u=𝒪(|x|β)u=\mathcal{O}(|x|^{\beta}) as |x||x|\to\infty.

2.2. A weighted positive mass theorem

For a weighted manifold (Mn,g,f)(M^{n},g,f), define the metric

(2.5) g~=e2fn1g.\tilde{g}=e^{-\frac{2f}{n-1}}g.

To prove Theorem 1.3, we need two lemmas which will enable a reduction to the unweighted positive mass theorem. Henceforth, denote by RR and R~\tilde{R} the scalar curvatures of gg and g~\tilde{g} respectively, and Rf=R+2Δf|f|2R_{f}=R+2\Delta f-|\nabla f|^{2} the weighted scalar curvature of (M,g,f)(M,g,f), where covariant derivatives are taken with respect to gg.

Lemma 2.4.

Let (Mn,g,f)(M^{n},g,f) be an AE weighted manifold of order τ>n22\tau>\frac{n-2}{2}, such that f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g). Then (Mn,g~)(M^{n},\tilde{g}) is an AE manifold of order τ\tau and

(2.6) R~=e2n1f(Rf+1n1|f|2).\tilde{R}=e^{\frac{2}{n-1}f}\left(R_{f}+\frac{1}{n-1}|\nabla f|^{2}\right).
Proof.

Since f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M), in asymptotic coordinates for the end MM_{\infty} of MM we have

(2.7) f=𝒪(ρτ),kf=𝒪(ρτ1),klf=𝒪(ρτ2).f=\mathcal{O}(\rho^{-\tau}),\quad\partial_{k}f=\mathcal{O}(\rho^{-\tau-1}),\quad\partial_{k}\partial_{l}f=\mathcal{O}(\rho^{-\tau-2}).

In particular,

(2.8) e2fn1=1+𝒪(ρτ)e^{-\frac{2f}{n-1}}=1+\mathcal{O}(\rho^{-\tau})

and so

(2.9) g~ij=e2fn1gij=(1+𝒪(ρτ))(δij+𝒪(ρτ))=δij+𝒪(ρτ).\tilde{g}_{ij}=e^{-\frac{2f}{n-1}}g_{ij}=(1+\mathcal{O}(\rho^{-\tau}))(\delta_{ij}+\mathcal{O}(\rho^{-\tau}))=\delta_{ij}+\mathcal{O}(\rho^{-\tau}).

Using (2.7), (2.8), and the asymptotic Euclideanness of gg, it also follows that

(2.10) kg~ij\displaystyle\partial_{k}\tilde{g}_{ij} =2n1e2fn1(kf)gij+e2fn1kgij=𝒪(ρτ1).\displaystyle=-\frac{2}{n-1}e^{-\frac{2f}{n-1}}(\partial_{k}f)g_{ij}+e^{-\frac{2f}{n-1}}\partial_{k}g_{ij}=\mathcal{O}(\rho^{-\tau-1}).

Similarly, klg~ij=𝒪(ρτ2)\partial_{k}\partial_{l}\tilde{g}_{ij}=\mathcal{O}(\rho^{-\tau-2}). These are the required decay conditions on g~\tilde{g}.

To prove (2.6), let φ=en22(n1)f\varphi=e^{-\frac{n-2}{2(n-1)}f}. We have g~=φ4n2g\tilde{g}=\varphi^{\frac{4}{n-2}}g and

(2.11) Δgφ=en22(n1)f(n22(n1)Δgf+(n22(n1))2|f|2).\Delta_{g}\varphi=e^{-\frac{n-2}{2(n-1)}f}\left(-\frac{n-2}{2(n-1)}\Delta_{g}f+\left(\frac{n-2}{2(n-1)}\right)^{2}|\nabla f|^{2}\right).

Recall that if gg is a Riemannian metric on an nn-dimensional manifold and φ\varphi is a smooth positive function, then the conformal metric φ4n2g\varphi^{\frac{4}{n-2}}g has scalar curvature

(2.12) φn+2n2(4(n1)n2Δgφ+Rφ).\varphi^{-\frac{n+2}{n-2}}\left(-\frac{4(n-1)}{n-2}\Delta_{g}\varphi+R\varphi\right).

Using this together with (2.11), it follows that

(2.13) R~\displaystyle\tilde{R} =en+22(n1)f(4(n1)n2en22(n1)f(n22(n1)Δgf+(n22(n1))2|f|2)+Ren22(n1)f)\displaystyle=e^{\frac{n+2}{2(n-1)}f}\left(-\frac{4(n-1)}{n-2}e^{-\frac{n-2}{2(n-1)}f}\left(-\frac{n-2}{2(n-1)}\Delta_{g}f+\left(\frac{n-2}{2(n-1)}\right)^{2}|\nabla f|^{2}\right)+Re^{-\frac{n-2}{2(n-1)}f}\right)
(2.14) =e2fn1(2Δgfn2n1|f|2+R)\displaystyle=e^{\frac{2f}{n-1}}\left(2\Delta_{g}f-\frac{n-2}{n-1}|\nabla f|^{2}+R\right)
(2.15) =e2fn1(Rf+1n1|f|2).\displaystyle=e^{\frac{2f}{n-1}}\left(R_{f}+\frac{1}{n-1}|\nabla f|^{2}\right).

Since RfL1(M,g)R_{f}\in L^{1}(M,g) by hypothesis, |f|2=𝒪(ρ2τ2)|\nabla f|^{2}=\mathcal{O}(\rho^{-2\tau-2}) by (2.7), and we have the decay (2.8), the formula (2.15) implies that R~L1(M,g)\tilde{R}\in L^{1}(M,g). As g~\tilde{g} and gg are asymptotically equivalent this implies R~L1(M,g~)\tilde{R}\in L^{1}(M,\tilde{g}). So g~\tilde{g} is an asymptotically Euclidean metric. ∎

The conformal choice (2.5) yields a nice formula for the weighted mass 𝔪f(g)\mathfrak{m}_{f}(g) defined in (1.2):

Lemma 2.5.

Let (Mn,g,f)(M^{n},g,f) be an AE weighted manifold of order τ>n22\tau>\frac{n-2}{2}. Assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g). Then the weighted mass of (M,g,f)(M,g,f) equals the unweighted mass of (M,g~)(M,\tilde{g}):

(2.16) 𝔪f(g)=𝔪(g~).\mathfrak{m}_{f}(g)=\mathfrak{m}(\tilde{g}).
Proof.

By Lemma 2.4, (M,g~)(M,\tilde{g}) is AE of order τ\tau so 𝔪(g~)\mathfrak{m}(\tilde{g}) is well-defined. Since g~=e2fn1g\tilde{g}=e^{-\frac{2f}{n-1}}g, we have

(2.17) 𝔪(g~)\displaystyle\mathfrak{m}(\tilde{g}) =limρSρi,j=1n(ig~ijjg~ii)𝐧jdA\displaystyle=\lim_{\rho\rightarrow\infty}\int_{S_{\rho}}\sum_{i,j=1}^{n}(\partial_{i}\tilde{g}_{ij}-\partial_{j}\tilde{g}_{ii})\mathbf{n}_{j}\,dA
(2.18) =limρSρe2fn1i,j=1n(igijjgii)𝐧jdA=:I12n1limρSρe2fn1i,j=1n((if)gij(jf)gii)𝐧jdA=:I2.\displaystyle=\underbrace{\lim_{\rho\to\infty}\int_{S_{\rho}}e^{-\frac{2f}{n-1}}\sum_{i,j=1}^{n}(\partial_{i}g_{ij}-\partial_{j}g_{ii})\mathbf{n}_{j}\,dA}_{=:I_{1}}-\underbrace{\frac{2}{n-1}\lim_{\rho\to\infty}\int_{S_{\rho}}e^{-\frac{2f}{n-1}}\sum_{i,j=1}^{n}((\partial_{i}f)g_{ij}-(\partial_{j}f)g_{ii})\mathbf{n}_{j}\,dA}_{=:I_{2}}.

Since e2fn1=1+𝒪(ρτ)e^{-\frac{2f}{n-1}}=1+\mathcal{O}(\rho^{-\tau}) and igij,jgii=𝒪(ρτ1)\partial_{i}g_{ij},\partial_{j}g_{ii}=\mathcal{O}(\rho^{-\tau-1}), we have

(2.19) I1\displaystyle I_{1} =limρSρi,j=1n(igijjgii)𝐧jdA+limρSρ𝒪(ρ2τ1)𝑑A=𝔪(g).\displaystyle=\lim_{\rho\to\infty}\int_{S_{\rho}}\sum_{i,j=1}^{n}(\partial_{i}g_{ij}-\partial_{j}g_{ii})\mathbf{n}_{j}\,dA+\lim_{\rho\to\infty}\int_{S_{\rho}}\mathcal{O}(\rho^{-2\tau-1})\,dA=\mathfrak{m}(g).

To handle I2I_{2}, use that gij=δij+𝒪(ρτ)g_{ij}=\delta_{ij}+\mathcal{O}(\rho^{-\tau}) to get

(2.20) I2\displaystyle I_{2} =2n1limρSρ(1+𝒪(ρτ))i,j=1n[(if)(δij+𝒪(ρτ))(jf)(1+𝒪(ρτ))]𝐧jdA\displaystyle=\frac{2}{n-1}\lim_{\rho\to\infty}\int_{S_{\rho}}(1+\mathcal{O}(\rho^{-\tau}))\sum_{i,j=1}^{n}[(\partial_{i}f)(\delta_{ij}+\mathcal{O}(\rho^{-\tau}))-(\partial_{j}f)(1+\mathcal{O}(\rho^{-\tau}))]\mathbf{n}_{j}\,dA
(2.21) =2n1limρSρ(1n)j=1n(jf)𝐧j+𝒪(ρ2τ1)dA\displaystyle=\frac{2}{n-1}\lim_{\rho\to\infty}\int_{S_{\rho}}(1-n)\sum_{j=1}^{n}(\partial_{j}f)\mathbf{n}_{j}+\mathcal{O}(\rho^{-2\tau-1})\,dA
(2.22) =2limρSρf,𝐧ef𝑑A.\displaystyle=-2\lim_{\rho\to\infty}\int_{S_{\rho}}\langle\nabla f,\mathbf{n}\rangle e^{-f}\,dA.

Using (2.19) and (2.22) in (2.18), we have

(2.23) 𝔪(g~)\displaystyle\mathfrak{m}(\tilde{g}) =I1I2=𝔪(g)+2limρSρf,𝐧ef𝑑A=𝔪f(g).\displaystyle=I_{1}-I_{2}=\mathfrak{m}(g)+2\lim_{\rho\to\infty}\int_{S_{\rho}}\langle\nabla f,\mathbf{n}\rangle e^{-f}\,dA=\mathfrak{m}_{f}(g).

We will now prove Theorem 1.3 using the previous two lemmas.

Proof of Theorem 1.3.

Let (Mn,g,f)(M^{n},g,f) be a weighted manifold which is AE of order τ>n22\tau>\frac{n-2}{2}, and suppose 3n73\leq n\leq 7 or MM is spin. Also assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g).

If Rf1n1|f|2R_{f}\geq-\frac{1}{n-1}|\nabla f|^{2}, then Lemma 2.4 shows that the conformal metric g~=e2fn1g\tilde{g}=e^{-\frac{2f}{n-1}}g has scalar curvature R~0\tilde{R}\geq 0. By Lemma 2.5 and Theorem 1.1,

(2.24) 𝔪f(g)=𝔪(g~)0.\mathfrak{m}_{f}(g)=\mathfrak{m}(\tilde{g})\geq 0.

This proves the first claim in part (a) of Theorem 1.3, as well as part (c).

Now suppose Rf0R_{f}\geq 0 and 𝔪f(g)=0\mathfrak{m}_{f}(g)=0. Lemmas 2.4 and 2.5 then imply R~0\tilde{R}\geq 0 and 𝔪(g~)=0\mathfrak{m}(\tilde{g})=0, so the rigidity part of Theorem 1.1 gives that (Mn,g~)(M^{n},\tilde{g}) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}). Thus

(2.25) 0=R~=e2n1f(Rf+1n1|f|2)1n1e2n1f|f|20.0=\tilde{R}=e^{\frac{2}{n-1}f}\left(R_{f}+\frac{1}{n-1}|\nabla f|^{2}\right)\geq\frac{1}{n-1}e^{\frac{2}{n-1}f}|\nabla f|^{2}\geq 0.

It follows that ff is constant, but since f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M), we have f0f\equiv 0. So g~=g\tilde{g}=g, and (Mn,g)(M^{n},g) too is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}). This proves the rigidity part of Theorem 1.3(a).

Theorem 1.3(a) implies Theorem 1.1 by taking the weight to be f0f\equiv 0. Conversely, we have just used Theorem 1.1 to prove Theorem 1.3(a). Hence Theorem 1.3(b) follows. ∎

2.3. The mass of a smooth metric measure space

In Definition 1.5, the mass of an AE SMMS =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m), mm\in\mathbb{R}, was defined as the weighted mass of (Mn,g,f)(M^{n},g,f). We will now motivate this by considering the mm\in\mathbb{N} case.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be a SMMS with mm\in\mathbb{N}, let (Fm,h)(F^{m},h) be an mm-dimensional scalar-flat manifold, and form the warped product

(Mn×Fm,g¯=ge2fmh).(M^{n}\times F^{m},\overline{g}=g\oplus e^{-\frac{2f}{m}}h).

Then the warped product has scalar curvature RfmR^{m}_{f}, and its Ricci tensor restricted to tangent vectors on MM is Ricfm\mathrm{Ric}^{m}_{f}. These are the curvatures associated to \mathcal{M}, defined in (1.6) and (1.7). We may therefore view the warped product as a natural ‘extrinsic’ space associated to \mathcal{M}. For this reason, if we are to define a mass for AE SMMSs which makes use of the parameter mm, it is natural to define it as the mass of the warped product.222Although the warped product is not AE, its mass can be defined using an integral similar to (2.2) as done in (2.26). We normalize by taking (Fm,h)(F^{m},h) to have unit volume.

Definition 2.6.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be an AE SMMS with mm\in\mathbb{N}. Let (Fm,h)(F^{m},h) be an mm-dimensional scalar-flat manifold of unit volume, and form the warped product (Mn×Fm,g¯=ge2fmh)(M^{n}\times F^{m},\overline{g}=g\oplus e^{-\frac{2f}{m}}h). Work in asymptotic coordinates for (Mn,g)(M^{n},g) and an orthonormal frame for (Fm,h)(F^{m},h). The mass of \mathcal{M} is defined as

(2.26) 𝔪():=limρSρM×F(ig¯ijjg¯aa)𝐧j𝑑A𝑑Volh,\mathfrak{m}(\mathcal{M}):=\lim_{\rho\to\infty}\int_{S^{M}_{\rho}\times F}(\partial_{i}\bar{g}_{ij}-\partial_{j}\bar{g}_{aa})\mathbf{n}_{j}\,dA\,d\mathrm{Vol}_{h},

where SρMS^{M}_{\rho} is the coordinate sphere of radius ρ\rho in asymptotic coordinates for MM, 𝐧j\mathbf{n}_{j} is the jj-th component of the outward Euclidean unit normal to SρMS^{M}_{\rho}, and dAdA is the Euclidean area element. Here the indices i,ji,j are summed over the coordinates for MM and the index aa is summed over the combined frame for M×FM\times F.

The next proposition reveals that (2.26) is nothing but the weighted mass of (Mn,g,f)(M^{n},g,f) (hence independent of both mm and FF). This is why we have used the weighted mass as the definition of mass for SMMSs even when mm\notin\mathbb{N}.

Proposition 2.7.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be an AE SMMS with mm\in\mathbb{N}. Then

(2.27) 𝔪()=𝔪f(g),\mathfrak{m}(\mathcal{M})=\mathfrak{m}_{f}(g),

where 𝔪f(g)\mathfrak{m}_{f}(g) is the weighted mass of (Mn,g,f)(M^{n},g,f) defined in (1.2).

Proof.

Fix asymptotic coordinates for MM, and take an orthonormal frame for FF so that hh is the m×mm\times m identity matrix. Putting these together yields a frame for the warped product (Mn×Fm,g¯=ge2fmh)(M^{n}\times F^{m},\overline{g}=g\oplus e^{-\frac{2f}{m}}h). According to Definition 2.6, the mass of \mathcal{M} is

(2.28) 𝔪()\displaystyle\mathfrak{m}(\mathcal{M}) =limρSρM×F(ig¯ijjg¯iijg¯ββ)𝐧j𝑑A𝑑Volh\displaystyle=\lim_{\rho\to\infty}\int_{S^{M}_{\rho}\times F}(\partial_{i}\bar{g}_{ij}-\partial_{j}\bar{g}_{ii}-\partial_{j}\bar{g}_{\beta\beta})\mathbf{n}_{j}\,dA\,d\mathrm{Vol}_{h}

where i,ji,j run over the coordinates for MM and β\beta runs over the coordinates for FF. Now we have

(2.29) g¯ij=gij,g¯ββ=e2fmhββ=e2fm,jg¯ββ=2me2fmjf.\bar{g}_{ij}=g_{ij},\quad\bar{g}_{\beta\beta}=e^{-\frac{2f}{m}}h_{\beta\beta}=e^{-\frac{2f}{m}},\quad\partial_{j}\bar{g}_{\beta\beta}=-\frac{2}{m}e^{-\frac{2f}{m}}\partial_{j}f.

Substituting these into (2.28) and using the fact that (Fm,h)(F^{m},h) has unit volume, we have

(2.30) 𝔪()\displaystyle\mathfrak{m}(\mathcal{M}) =limρSρM×F(igijjgii)𝐧j𝑑A𝑑Volh+limρSρM×F2e2fm(jf)𝐧j𝑑A𝑑Volh\displaystyle=\lim_{\rho\to\infty}\int_{S^{M}_{\rho}\times F}(\partial_{i}g_{ij}-\partial_{j}g_{ii})\mathbf{n}_{j}\,dA\,d\mathrm{Vol}_{h}+\lim_{\rho\to\infty}\int_{S^{M}_{\rho}\times F}2e^{-\frac{2f}{m}}(\partial_{j}f)\mathbf{n}_{j}\,dA\,d\mathrm{Vol}_{h}
(2.31) =(𝔪(g)+2limρSρMf,𝐧ef𝑑A)Vol(Fm,h)\displaystyle=\left(\mathfrak{m}(g)+2\lim_{\rho\to\infty}\int_{S^{M}_{\rho}}\langle\nabla f,\mathbf{n}\rangle e^{-f}\,dA\right)\mathrm{Vol}(F^{m},h)
(2.32) =𝔪f(g).\displaystyle=\mathfrak{m}_{f}(g).

2.4. A positive mass theorem for smooth metric measure spaces

Proof of Theorem 1.6.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be an AE SMMS of order τ>n22\tau>\frac{n-2}{2} with m(1n,0]m\in\mathbb{R}\setminus(1-n,0], and suppose 3n73\leq n\leq 7 or MM is spin. Also assume that

f𝒞τ2,α(M),RfL1(M,g),Rfm0.f\in\mathcal{C}^{2,\alpha}_{-\tau}(M),\quad R_{f}\in L^{1}(M,g),\quad R^{m}_{f}\geq 0.

Since m(1n,0]m\notin(1-n,0], we have

(2.33) 0Rfm=Rf1m|f|2Rf+1n1|f|2.0\leq R^{m}_{f}=R_{f}-\frac{1}{m}|\nabla f|^{2}\leq R_{f}+\frac{1}{n-1}|\nabla f|^{2}.

It follows by Theorem 1.3(c) that 𝔪()=𝔪f(g)0\mathfrak{m}(\mathcal{M})=\mathfrak{m}_{f}(g)\geq 0. If equality holds, then using the conformal metric g~=e2fn1g\tilde{g}=e^{-\frac{2f}{n-1}}g and Lemma 2.5, we have

(2.34) 𝔪()=𝔪f(g)=𝔪(g~)=0.\mathfrak{m}(\mathcal{M})=\mathfrak{m}_{f}(g)=\mathfrak{m}(\tilde{g})=0.

By Lemma 2.4, the scalar curvature of g~\tilde{g} is

(2.35) R~\displaystyle\tilde{R} =e2n1f(Rf+1n1|f|2)=e2n1f(Rfm+(1m+1n1)|f|2).\displaystyle=e^{\frac{2}{n-1}f}\left(R_{f}+\frac{1}{n-1}|\nabla f|^{2}\right)=e^{\frac{2}{n-1}f}\left(R^{m}_{f}+\left(\frac{1}{m}+\frac{1}{n-1}\right)|\nabla f|^{2}\right).

Combining the first equality with (2.33) yields R~0\tilde{R}\geq 0, so by (2.34) and the rigidity in the unweighted positive mass theorem, we have (M,g~)(n,δij)(M,\tilde{g})\cong(\mathbb{R}^{n},\delta_{ij}). Now (2.35) gives

(2.36) 0=R~=e2n1f(Rfm+(1m+1n1)|f|2)e2n1f(1m+1n1)|f|20,0=\tilde{R}=e^{\frac{2}{n-1}f}\left(R^{m}_{f}+\left(\frac{1}{m}+\frac{1}{n-1}\right)|\nabla f|^{2}\right)\geq e^{\frac{2}{n-1}f}\left(\frac{1}{m}+\frac{1}{n-1}\right)|\nabla f|^{2}\geq 0,

which implies f=0\nabla f=0. But f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M), so f0f\equiv 0. Thus g~=g\tilde{g}=g, so (Mn,g)(n,δij)(M^{n},g)\cong(\mathbb{R}^{n},\delta_{ij}). ∎

3. Spin geometry on warped products

We now turn to spinorial aspects of SMMSs. We will restrict to SMMSs =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) with mm\in\mathbb{N}, in which case \mathcal{M} is closely associated to a warped product manifold

(3.1) (Mn×Fm,g¯=ge2fmh)(M^{n}\times F^{m},\bar{g}=g\oplus e^{-\frac{2f}{m}}h)

where (Fm,h)(F^{m},h) is scalar-flat (see §1 or §2.3). In this section, MM and FF are assumed to be spin manifolds.

We study the spin geometry of \mathcal{M} via the spin geometry of the warped product (M×F,g¯)(M\times F,\bar{g}). Doing this entails relating spinors on the warped product to spinors on the factors (M,g)(M,g) and (F,h)(F,h). Facilitated by [32, §3.2], we carry this out in §3.2 by identifying the appropriate spinor bundles, and in §3.3 by relating their Dirac operators (Theorem 3.5). We begin by reviewing the necessary spin geometry; for comprehensive treatments see [9, 20, 23].

3.1. Generalities on spin geometry

Let EXE\to X be a oriented vector bundle of rank kk equipped with a metric hh, and let PSOEXP_{\mathrm{SO}}E\to X be the bundle of positive orthonormal bases of (E,h)(E,h). Recall that a spin structure on EE is a principal Spink\mathrm{Spin}_{k}-bundle PSpinEXP_{\mathrm{Spin}}E\to X with an equivariant double cover PSpinEPSOEP_{\mathrm{Spin}}E\to P_{\mathrm{SO}}E with respect to the right group actions and the universal covering SpinkSOk\mathrm{Spin}_{k}\to\mathrm{SO}_{k}. If PSpinEP_{\mathrm{Spin}}E is a spin structure for (E,h)(E,h), we can form the spinor bundle ΣE=PSpinE×ρ2k/2\Sigma E=P_{\mathrm{Spin}}E\times_{\rho}\mathbb{C}^{2^{\lfloor k/2\rfloor}} over XX, where ρ\rho is the restriction to Spink\mathrm{Spin}_{k} of an irreducible representation of the complexified Clifford algebra l(k)\mathbb{C}\mathrm{l}(\mathbb{R}^{k}). Then ΣE\Sigma E has complex rank 2k22^{\lfloor\frac{k}{2}\rfloor}, and is a bundle of modules over the bundle of Clifford algebras l(E,h)\mathbb{C}\mathrm{l}(E,h).

Any l(E,h)\mathbb{C}\mathrm{l}(E,h)-module SS, such as ΣE\Sigma E, has a Hermitian metric such that the action of unit vectors is unitary. This Hermitian metric on SS is obtained by an averaging procedure which we now spell out. Take an arbitrary Hermitian metric (,)(\cdot,\cdot), and for each xXx\in X let Γx\Gamma_{x} be the finite subgroup of l(Ex,hx)\mathbb{C}\mathrm{l}(E_{x},h_{x}) generated by an orthonormal basis for ExE_{x}. Now define the Hermitian metric ,\langle\cdot,\cdot\rangle on SS by setting for all ψ1,ψ2Sx\psi_{1},\psi_{2}\in S_{x}

(3.2) ψ1,ψ2=1|Γx|τΓx(τψ1,τψ2).\langle\psi_{1},\psi_{2}\rangle=\frac{1}{|\Gamma_{x}|}\sum_{\tau\in\Gamma_{x}}(\tau\cdot\psi_{1},\tau\cdot\psi_{2}).

This averaged metric is the Hermitian metric on SS, and is unique up to positive scaling. Note that if we repeat the above using ,\langle\cdot,\cdot\rangle as the starting metric, then the averaging procedure recovers ,\langle\cdot,\cdot\rangle.

A Riemannian manifold is spin if it admits a spin structure, meaning a spin structure on its tangent bundle. Given a spin structure PSpin(TX)P_{\mathrm{Spin}}(TX) on a spin manifold (Xn,g)(X^{n},g), the spinor bundle ΣX:=ΣTX\Sigma X:=\Sigma TX is a rank 2n22^{\lfloor\frac{n}{2}\rfloor} bundle of complex modules over l(TX,g)\mathbb{C}\mathrm{l}(TX,g). Moreover, ΣX\Sigma X gets a Hermitian metric using the averaging procedure described above (so that Clifford multiplication by unit tangent vectors is unitary), as well as a connection defined by

(3.3) Yψ=dψ(Y)+14j,k=1ng(Yej,ek)ejekψ,YTM,ψΓ(ΣX),\nabla_{Y}\psi=d\psi(Y)+\frac{1}{4}\sum_{j,k=1}^{n}g(\nabla_{Y}e_{j},e_{k})e_{j}\cdot e_{k}\cdot\psi,\quad Y\in TM,\psi\in\Gamma(\Sigma X),

where e1,,ene_{1},\ldots,e_{n} is an orthonormal basis for (TX,g)(TX,g). The Dirac operator D:Γ(ΣX)Γ(ΣX)D:\Gamma(\Sigma X)\to\Gamma(\Sigma X) is a symmetric first-order elliptic operator defined by

(3.4) Dψ=i=1neieiψ.D\psi=\sum_{i=1}^{n}e_{i}\cdot\nabla_{e_{i}}\psi.

If RR denotes the scalar curvature of (X,g)(X,g), then DD satisfies the Lichnerowicz formula

(3.5) D2ψ=Δψ+R4ψ,D^{2}\psi=-\Delta\psi+\frac{R}{4}\psi,

where Δ==eieieiei\Delta=-\nabla^{*}\nabla=\nabla_{e_{i}}\nabla_{e_{i}}-\nabla_{\nabla_{e_{i}}e_{i}} is the Laplacian on the spinor bundle.

Define a section ωX\omega_{X} of l(TX,g)\mathbb{C}\mathrm{l}(TX,g) as follows: if e1,,ene_{1},\ldots,e_{n} is a positive orthonormal basis for (TxX,g)(T_{x}X,g), then

(3.6) ωX(x)=in+12e1en.\omega_{X}(x)=i^{\lfloor\frac{n+1}{2}\rfloor}e_{1}\cdots e_{n}.

If n=dimXn=\dim X is even, then ωX2=1\omega_{X}^{2}=1 so ΣX\Sigma X splits as an orthogonal sum of the ±1\pm 1-eigenbundles:

(3.7) ΣX=Σ^+XΣ^X,Σ^±X={ψΣXωXψ=±ψ}.\Sigma X=\hat{\Sigma}^{+}X\oplus\hat{\Sigma}^{-}X,\quad\hat{\Sigma}^{\pm}X=\{\psi\in\Sigma X\mid\omega_{X}\cdot\psi=\pm\psi\}.

We write ψ=ψ++ψ\psi=\psi^{+}+\psi^{-} for the corresponding decomposition of ψΣX\psi\in\Sigma X. The conjugate spinor is defined by ψ¯=ψ+ψ\bar{\psi}=\psi^{+}-\psi^{-}. We have vωX=ωXvv\cdot\omega_{X}=-\omega_{X}\cdot v for all vTXv\in TX, so Clifford multiplication by vv permutes Σ^±X\hat{\Sigma}^{\pm}X.

3.2. The spinor bundle of a warped product

Let (Mn,g)(M^{n},g) and (Fm,h)(F^{m},h) be spin manifolds, fix a smooth function f𝒞(M)f\in\mathcal{C}^{\infty}(M), and consider the warped product

(3.8) (M×F,g¯=ge2fmh)(M\times F,\bar{g}=g\oplus e^{-\frac{2f}{m}}h)

along with the projection maps

(3.9) π1:M×FM,π2:M×FF.\pi_{1}:M\times F\to M,\quad\pi_{2}:M\times F\to F.

Fixing spin structures on MM and FF (that is, on (TM,g)(TM,g) and (TF,h)(TF,h) respectively), we describe the induced spin structure on (T(M×F),g¯)(T(M\times F),\bar{g}). Pulling back the chosen spin structures via π1\pi_{1} and π2\pi_{2} gives spin structures on the bundles (π1TM,π1g)(\pi_{1}^{*}TM,\pi_{1}^{*}g) and (π2TF,π2h)(\pi_{2}^{*}TF,\pi_{2}^{*}h). The latter determines a spin structure for the bundle (π2TF,e2fmπ2h)(\pi_{2}^{*}TF,e^{-\frac{2f}{m}}\pi_{2}^{*}h) which is hereafter called VV [23, §2, Remark 1.9]. Since

(3.10) (T(M×F),g¯)(π1TM,π1g)V,(T(M\times F),\bar{g})\cong(\pi_{1}^{*}TM,\pi_{1}^{*}g)\oplus V,

and the two summands on the right are now endowed with spin structures, their direct sum also gets a spin structure [23, §2, Proposition 1.15]. This is the induced spin structure on (T(M×F),g¯)(T(M\times F),\bar{g}).

The spin structures on (TM,g)(TM,g), (TF,h)(TF,h), V=(π2TF,e2fmπ2h)V=(\pi_{2}^{*}TF,e^{-\frac{2f}{m}}\pi_{2}^{*}h) and (T(M×F),g¯)(T(M\times F),\bar{g}) induce the spinor bundles ΣM\Sigma M, ΣF\Sigma F, ΣV\Sigma V and Σ¯(M×F)\bar{\Sigma}(M\times F) respectively, as listed in Table 1.

Spinor bundle Base space Module over…
ΣM\Sigma M MM l(TM,g)\mathbb{C}\mathrm{l}(TM,g)
ΣF\Sigma F FF l(TF,h)\mathbb{C}\mathrm{l}(TF,h)
ΣV\Sigma V M×FM\times F l(π2TF,e2fmπ2h)\mathbb{C}\mathrm{l}(\pi_{2}^{*}TF,e^{-\frac{2f}{m}}\pi_{2}^{*}h)
Σ¯(M×F)\bar{\Sigma}(M\times F) M×FM\times F l(T(M×F),g¯)\mathbb{C}\mathrm{l}(T(M\times F),\bar{g})
Table 1. The spinor bundles we will use.

We will describe some relationships between these bundles and the structures on them. Firstly, since ΣF\Sigma F and π2ΣF\pi_{2}^{*}\Sigma F are Clifford modules over l(TF,h)\mathbb{C}\mathrm{l}(TF,h) and l(π2TF,π2h)\mathbb{C}\mathrm{l}(\pi_{2}^{*}TF,\pi_{2}^{*}h) respectively, the averaging procedure described in §3.1 applies to give Hermitian metrics ,ΣF\langle\cdot,\cdot\rangle_{\Sigma F} and ,π2ΣF\langle\cdot,\cdot\rangle_{\pi_{2}^{*}\Sigma F}. It is not hard to see that ,π2ΣF=π2,ΣF\langle\cdot,\cdot\rangle_{\pi_{2}^{*}\Sigma F}=\pi_{2}^{*}\langle\cdot,\cdot\rangle_{\Sigma F}. The averaging procedure can also be used to define ,ΣV\langle\cdot,\cdot\rangle_{\Sigma V} on ΣV\Sigma V; on the other hand, we have the following useful identification of ΣV\Sigma V together with its inner product.

Lemma 3.1.

There is a bundle isometry (ΣV,,ΣV)(π2ΣF,,π2ΣF)(\Sigma V,\langle\cdot,\cdot\rangle_{\Sigma V})\cong(\pi_{2}^{*}\Sigma F,\langle\cdot,\cdot\rangle_{\pi_{2}^{*}\Sigma F}). If \cdot is the Clifford multiplication on π2ΣF\pi_{2}^{*}\Sigma F, then the Clifford multiplication ^\hat{\,\cdot\,} on ΣV\Sigma V is given under the identification ΣVπ2ΣF\Sigma V\cong\pi_{2}^{*}\Sigma F by

(3.11) π2TFπ2ΣFπ2ΣF,(v,ν)v^ν=efmvν.\pi_{2}^{*}TF\otimes\pi_{2}^{*}\Sigma F\to\pi_{2}^{*}\Sigma F,\quad(v,\nu)\mapsto v\hat{\,\cdot\,}\nu=e^{-\frac{f}{m}}v\cdot\nu.
Proof.

This is because VV is obtained by multiplying the bundle metric on (π2TF,π2h)(\pi_{2}^{*}TF,\pi_{2}^{*}h) by the conformal factor e2fme^{-\frac{2f}{m}} (see e.g. [9, pp.69], whose discussion generalizes to spin structures on vector bundles). ∎

We will use “\cdot” to denote the Clifford multiplication on ΣM\Sigma M, ΣF\Sigma F, Σ¯(M×F)\bar{\Sigma}(M\times F), and their pullbacks. In light of Lemma 3.1, ΣV\Sigma V will be understood to be the vector bundle π2ΣF\pi_{2}^{*}\Sigma F equipped with the Clifford multiplication “^\hat{\,\cdot\,}” given by (3.11). The next proposition identifies Σ¯(M×F)\bar{\Sigma}(M\times F) and its Clifford module structure.

Proposition 3.2.

We have

(3.12) Σ¯(M×F)π1(𝚺M)ΣV\bar{\Sigma}(M\times F)\cong\pi_{1}^{*}({}^{\diamond}\mathbf{\Sigma}M)\otimes\Sigma V

as vector bundles, where

(3.13) 𝚺M={ΣMif n or m is even,ΣMΣMif n and m are odd.{}^{\diamond}\mathbf{\Sigma}M=\begin{cases}\Sigma M&\text{if }n\text{ or }m\text{ is even},\\ \Sigma M\oplus\Sigma M&\text{if }n\text{ and }m\text{ are odd}.\end{cases}

Under this identification, the structure of Σ¯(M×F)\bar{\Sigma}(M\times F) as a module over l(T(M×F),g¯)\mathbb{C}\mathrm{l}(T(M\times F),\bar{g}) is given as follows. Let ϕπ1(𝚺M)\phi\in\pi_{1}^{*}{}^{\diamond}(\mathbf{\Sigma}M) and νΣV\nu\in\Sigma V. The Clifford multiplication of ϕνΣ¯(M×F)\phi\otimes\nu\in\bar{\Sigma}(M\times F) by (x,v)TMTF(x,v)\in TM\oplus TF is

(3.14) (x,v)(ϕν)={(xϕ)ν+ϕ¯(v^ν)n even(xϕ)ν¯+ϕ(v^ν)n odd, m even(xϕ1xϕ2)ν+(ϕ2ϕ1)(v^ν)n,m odd.(x,v)\cdot(\phi\otimes\nu)=\begin{cases}(x\cdot\phi)\otimes\nu+\bar{\phi}\otimes(v\hat{\,\cdot\,}\nu)&n\text{ even}\\ (x\cdot\phi)\otimes\bar{\nu}+\phi\otimes(v\hat{\,\cdot\,}\nu)&n\text{ odd, }m\text{ even}\\ (x\cdot\phi_{1}\oplus-x\cdot\phi_{2})\otimes\nu+(\phi_{2}\oplus\phi_{1})\otimes(v\hat{\,\cdot\,}\nu)&n,m\text{ odd}.\end{cases}

For the last case we have written ϕ=ϕ1+ϕ2ΣMΣM\phi=\phi_{1}+\phi_{2}\in\Sigma M\oplus\Sigma M.

Proof.

Since π1:(M×F,g¯)(M,g)\pi_{1}:(M\times F,\bar{g})\to(M,g) is clearly a Riemannian submersion, the proposition follows from the discussion in [32, §3.2] (see specifically equations (10), (11) and Notation 1 there). ∎

One verifies that, as should be the case, applying Clifford multiplication by (x,v)(x,v) twice has the same effect as scaling by g¯((x,v),(x,v))-\bar{g}((x,v),(x,v)).

The Hermitian metric on Σ¯(M×F)\bar{\Sigma}(M\times F) is again obtained by the averaging procedure from §3.1. Namely, using the identification (3.12), define an initial Hermitian metric on Σ¯(M×F)\bar{\Sigma}(M\times F) by

(3.15) (ϕν,ϕν)Σ¯(M×F)=ϕ,ϕπ1(𝚺M)ν,νΣV(\phi\otimes\nu,\phi^{\prime}\otimes\nu^{\prime})_{\bar{\Sigma}(M\times F)}=\langle\phi,\phi^{\prime}\rangle_{\pi_{1}^{*}({}^{\diamond}\mathbf{\Sigma}M)}\langle\nu,\nu^{\prime}\rangle_{\Sigma V}

and extending linearly. For each (x,y)M×F(x,y)\in M\times F, let Γx,y\Gamma_{x,y} be the finite Clifford subgroup generated by an orthonormal basis for (T(x,y)(M×F),g¯)(T_{(x,y)}(M\times F),\bar{g}) of the form

(3.16) {e1,,en,ε1,,εm}\{e_{1},\ldots,e_{n},\varepsilon_{1},\ldots,\varepsilon_{m}\}

where the eie_{i} are orthonormal for (TxM,g)(T_{x}M,g) and the εj\varepsilon_{j} are orthonormal for (TyF,e2f(x)mh)(T_{y}F,e^{-\frac{2f(x)}{m}}h). The final Hermitian metric on Σ¯(M×F)\bar{\Sigma}(M\times F) is then given by averaging as in (3.2):

(3.17) ϕν,ϕνΣ¯(M×F)=1|Γx,y|τΓx,y(τ(ϕν),τ(ϕν))Σ¯(M×F).\langle\phi\otimes\nu,\phi^{\prime}\otimes\nu^{\prime}\rangle_{\bar{\Sigma}(M\times F)}=\frac{1}{|\Gamma_{x,y}|}\sum_{\tau\in\Gamma_{x,y}}(\tau\cdot(\phi\otimes\nu),\tau\cdot(\phi^{\prime}\otimes\nu^{\prime}))_{\bar{\Sigma}(M\times F)}.

We leave it to the reader to check that ,Σ¯(M×F)\langle\cdot,\cdot\rangle_{\bar{\Sigma}(M\times F)} actually coincides with (,)Σ¯(M×F)(\cdot,\cdot)_{\bar{\Sigma}(M\times F)}. (This follows from the fact that ,π1(𝚺M)\langle\cdot,\cdot\rangle_{\pi_{1}^{*}({}^{\diamond}\mathbf{\Sigma}M)} and ,ΣV\langle\cdot,\cdot\rangle_{\Sigma V} are obtained from the averaging procedure, and averaging such a metric does not yield a different metric.) Thus we have:

Proposition 3.3.

The Hermitian metric on Σ¯(M×F)=π1(𝚺M)ΣV\bar{\Sigma}(M\times F)=\pi_{1}^{*}({}^{\diamond}\mathbf{\Sigma}M)\otimes\Sigma V is given by

(3.18) ϕν,ϕνΣ¯(M×F)=ϕ,ϕπ1(𝚺M)ν,νΣV\langle\phi\otimes\nu,\phi^{\prime}\otimes\nu^{\prime}\rangle_{\bar{\Sigma}(M\times F)}=\langle\phi,\phi^{\prime}\rangle_{\pi_{1}^{*}({}^{\diamond}\mathbf{\Sigma}M)}\langle\nu,\nu^{\prime}\rangle_{\Sigma V}

on decomposable spinors, and extending linearly.

3.3. The warped product spin connection and Dirac operator

In this subsection, we compute the connection and Dirac operator on Σ¯(M×F)\bar{\Sigma}(M\times F) by adapting the discussion in [32]. Recalling the orthogonal decomposition (T(M×F),g¯)(π1TM,π1g)V(T(M\times F),\bar{g})\cong(\pi_{1}^{*}TM,\pi_{1}^{*}g)\oplus V, define the projections

(3.19) ()H:T(M×F)π1TM,()V:T(M×F)V.(\cdot)^{H}:T(M\times F)\to\pi_{1}^{*}TM,\quad(\cdot)^{V}:T(M\times F)\to V.

Let ¯\bar{\nabla} be the Levi-Civita connection on (T(M×F),g¯)(T(M\times F),\bar{g}). Define the 2-tensors T,AT,A which act on X,YT(M×F)X,Y\in T(M\times F) by

(3.20) T(X,Y)\displaystyle T(X,Y) =(¯XVYV)H+(¯XVYH)V,\displaystyle=(\bar{\nabla}_{X^{V}}Y^{V})^{H}+(\bar{\nabla}_{X^{V}}Y^{H})^{V},
(3.21) A(X,Y)\displaystyle A(X,Y) =(¯XHYV)H+(¯XHYH)V.\displaystyle=(\bar{\nabla}_{X^{H}}Y^{V})^{H}+(\bar{\nabla}_{X^{H}}Y^{H})^{V}.

These were originally introduced by O’Neill [29] to study the curvatures of a Riemannian submersion.

We now choose convenient local frames for (T(M×F),g¯)(T(M\times F),\bar{g}). Let (ξ1,,ξn)(\xi_{1},\ldots,\xi_{n}) be a local orthonormal frame for (TM,g)(TM,g). Pulling this back by π1\pi_{1} gives nn local orthonormal vector fields on T(M×F)T(M\times F) which we also call ξ1,,ξn\xi_{1},\ldots,\xi_{n}. Let (η1,,ηm)(\eta_{1},\ldots,\eta_{m}) be a local orthonormal frame for (TF,h)(TF,h). Pulling this back by π2\pi_{2} gives mm local vector fields on T(M×F)T(M\times F) which we also call η1,,ηm\eta_{1},\ldots,\eta_{m}. Define ζi=efmηi\zeta_{i}=e^{\frac{f}{m}}\eta_{i} for i=1,,mi=1,\ldots,m. Then

(ξ1,,ξn,ζ1,,ζm)(\xi_{1},\ldots,\xi_{n},\zeta_{1},\ldots,\zeta_{m})

is a local orthonormal frame for (T(M×F),g¯)(T(M\times F),\bar{g}). We will also assume that (ξ1,,ξn)(\xi_{1},\ldots,\xi_{n}) coincides with normal coordinates for (M,g)(M,g) at a point x0x_{0}, and that (η1,,ηm)(\eta_{1},\ldots,\eta_{m}) coincides with normal coordinates for (F,h)(F,h) at y0y_{0}. Using that g¯=ge2fmh\bar{g}=g\oplus e^{-\frac{2f}{m}}h, routine computations yield the following identities at (x0,y0)(x_{0},y_{0}):

(3.22) ¯ξαξβ=¯ξαζi=0,¯ζiζj=1mδijf,g¯(¯ζiζj,ζk)=0,\displaystyle\bar{\nabla}_{\xi_{\alpha}}\xi_{\beta}=\bar{\nabla}_{\xi_{\alpha}}\zeta_{i}=0,\quad\bar{\nabla}_{\zeta_{i}}\zeta_{j}=\frac{1}{m}\delta_{ij}\nabla f,\quad\bar{g}(\bar{\nabla}_{\zeta_{i}}\zeta_{j},\zeta_{k})=0,
(3.23) A(ξα,ξβ)=A(ξα,ζi)=0,T(ζi,ζj)=1mδijf.\displaystyle A(\xi_{\alpha},\xi_{\beta})=A(\xi_{\alpha},\zeta_{i})=0,\quad T(\zeta_{i},\zeta_{j})=\frac{1}{m}\delta_{ij}\nabla f.
Definition 3.4.

We call the frame (ξ1,,ξn,ζ1,,ζm)(\xi_{1},\ldots,\xi_{n},\zeta_{1},\ldots,\zeta_{m}) constructed above a split orthonormal frame for (T(M×F),g¯)(T(M\times F),\bar{g}) centered at the point (x0,y0)M×F(x_{0},y_{0})\in M\times F. Thus the identities (3.22), (3.23) hold at (x0,y0)(x_{0},y_{0}).

The next theorem computes the connection and Dirac operator on Σ¯(M×F)\bar{\Sigma}(M\times F) by describing how they behave on spinors which are decomposable with respect to the identification (3.12). This actually computes the full connection and Dirac operator, because Σ¯(M×F)\bar{\Sigma}(M\times F) is locally trivialized by decomposable spinors, and the connection and Dirac operator obey Leibniz-type rules. For the Dirac operator, this is

(3.24) D¯(uψ)=uD¯ψ+¯uψ,u𝒞(M×F),ψΓ(Σ¯(M×F)).\bar{D}(u\psi)=u\bar{D}\psi+\bar{\nabla}u\cdot\psi,\quad u\in\mathcal{C}^{\infty}(M\times F),\quad\psi\in\Gamma(\bar{\Sigma}(M\times F)).
Theorem 3.5.

Let ψΓ(Σ¯(M×F))\psi\in\Gamma(\bar{\Sigma}(M\times F)) be a spinor of the form ψ=π1ϕπ2ν\psi=\pi_{1}^{*}\phi\otimes\pi_{2}^{*}\nu, where ϕΓ(𝚺M)\phi\in\Gamma({}^{\diamond}\mathbf{\Sigma}M) and νΓ(ΣF)\nu\in\Gamma(\Sigma F). For simplicity, identify ϕ\phi with π1ϕ\pi_{1}^{*}\phi and ν\nu with π2ν\pi_{2}^{*}\nu. Write ¯\bar{\nabla} and D¯\bar{D} for the connection and Dirac operator, respectively, on Σ¯(M×F)\bar{\Sigma}(M\times F). For all XTMX\in TM, YTFY\in TF we have

(3.25) ¯Xψ\displaystyle\bar{\nabla}_{X}\psi =Xϕν,\displaystyle=\nabla_{X}\phi\otimes\nu,
(3.26) ¯Yψ\displaystyle\bar{\nabla}_{Y}\psi =ϕYFν+12mYfψ,\displaystyle=\phi\otimes\nabla^{F}_{Y}\nu+\frac{1}{2m}Y\cdot\nabla f\cdot\psi,
(3.27) D¯ψ\displaystyle\bar{D}\psi ={Dfϕν+efmϕ¯DFνn evenDfϕν¯+efmϕDFνn odd, m even(Dfϕ1Dfϕ2)ν+efm(ϕ2ϕ1)DFνn,m odd,\displaystyle=\begin{cases}D_{f}\phi\otimes\nu+e^{\frac{f}{m}}\bar{\phi}\otimes D^{F}\nu&n\text{ even}\\ D_{f}\phi\otimes\bar{\nu}+e^{\frac{f}{m}}\phi\otimes D^{F}\nu&n\text{ odd, }m\text{ even}\\ (D_{f}\phi_{1}\oplus-D_{f}\phi_{2})\otimes\nu+e^{\frac{f}{m}}(\phi_{2}\oplus\phi_{1})\otimes D^{F}\nu&n,m\text{ odd},\end{cases}

where Df=D12fD_{f}=D-\frac{1}{2}\nabla f\cdot is the weighted Dirac operator on ΣM\Sigma M, and DFD^{F} is the Dirac operator on ΣF\Sigma F.

Proof.

Let (ξ1,,ξn,ζ1,,ζm)(\xi_{1},\ldots,\xi_{n},\zeta_{1},\ldots,\zeta_{m}) be a split orthonormal frame for (T(M×F),g¯)(T(M\times F),\bar{g}) centered at (x0,y0)M×F(x_{0},y_{0})\in M\times F, in the sense of Definition 3.4. Since π1:(M×F,g¯)(M,g)\pi_{1}:(M\times F,\bar{g})\to(M,g) is a Riemannian submersion, we can apply [32, Lemma 6] to get

(3.28) ¯ξαψ\displaystyle\bar{\nabla}_{\xi_{\alpha}}\psi =ξα𝒯ψ+12β=1nξβA(ξα,ξβ)ψ,\displaystyle=\nabla_{\xi_{\alpha}}^{\mathcal{T}}\psi+\frac{1}{2}\sum_{\beta=1}^{n}\xi_{\beta}\cdot A(\xi_{\alpha},\xi_{\beta})\cdot\psi,
(3.29) ¯ζiψ\displaystyle\bar{\nabla}_{\zeta_{i}}\psi =ζiZψ+12j=1mζjT(ζi,ζj)ψ+14α=1nξαA(ξα,ζi)ψ,\displaystyle=\nabla_{\zeta_{i}}^{Z}\psi+\frac{1}{2}\sum_{j=1}^{m}\zeta_{j}\cdot T(\zeta_{i},\zeta_{j})\cdot\psi+\frac{1}{4}\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot A(\xi_{\alpha},\zeta_{i})\cdot\psi,

where

(3.30) ξα𝒯(ϕν)\displaystyle\nabla_{\xi_{\alpha}}^{\mathcal{T}}(\phi\otimes\nu) :=ξαϕν+ϕξα𝒱ν,\displaystyle:=\nabla_{\xi_{\alpha}}\phi\otimes\nu+\phi\otimes\nabla_{\xi_{\alpha}}^{\mathcal{V}}\nu,\qquad ξα𝒱ν:=dν(ξα)+14j,k=1mg¯(¯ξαζj,ζk)ζj^ζk^ν,\displaystyle\nabla^{\mathcal{V}}_{\xi_{\alpha}}\nu:=d\nu(\xi_{\alpha})+\frac{1}{4}\sum_{j,k=1}^{m}\bar{g}(\bar{\nabla}_{\xi_{\alpha}}\zeta_{j},\zeta_{k})\zeta_{j}\hat{\,\cdot\,}\zeta_{k}\hat{\,\cdot\,}\nu,
(3.31) ζiZ(ϕν)\displaystyle\nabla_{\zeta_{i}}^{Z}(\phi\otimes\nu) :=ϕζiZν,\displaystyle:=\phi\otimes\nabla_{\zeta_{i}}^{Z}\nu, ζiZν:=dν(ζi)+14j,k=1mg¯(¯ζiζj,ζk)ζj^ζk^ν.\displaystyle\nabla_{\zeta_{i}}^{Z}\nu:=d\nu(\zeta_{i})+\frac{1}{4}\sum_{j,k=1}^{m}\bar{g}(\bar{\nabla}_{\zeta_{i}}\zeta_{j},\zeta_{k})\zeta_{j}\hat{\,\cdot\,}\zeta_{k}\hat{\,\cdot\,}\nu.

All computations in the rest of this proof are done at the point (x0,y0)(x_{0},y_{0}) where the identities (3.22), (3.23) hold. Since ¯ξαζj=0\bar{\nabla}_{\xi_{\alpha}}\zeta_{j}=0 and ν\nu is constant in the MM directions, we have ξα𝒱ν=0\nabla_{\xi_{\alpha}}^{\mathcal{V}}\nu=0. Combined with the fact that A(ξα,ξβ)=0A(\xi_{\alpha},\xi_{\beta})=0, (3.28) becomes

(3.32) ¯ξαψ\displaystyle\bar{\nabla}_{\xi_{\alpha}}\psi =ξαϕν,\displaystyle=\nabla_{\xi_{\alpha}}\phi\otimes\nu,

which implies (3.25). Since T(ζi,ζj)=1mδijfT(\zeta_{i},\zeta_{j})=\frac{1}{m}\delta_{ij}\nabla f and A(ξα,ζi)=0A(\xi_{\alpha},\zeta_{i})=0, (3.29) becomes

(3.33) ¯ζiψ\displaystyle\bar{\nabla}_{\zeta_{i}}\psi =ϕζiZν+12mζifψ.\displaystyle=\phi\otimes\nabla^{Z}_{\zeta_{i}}\nu+\frac{1}{2m}\zeta_{i}\cdot\nabla f\cdot\psi.

We have g¯(¯ζiζj,ζk)=0\bar{g}(\bar{\nabla}_{\zeta_{i}}\zeta_{j},\zeta_{k})=0; also, since the local orthonormal frame (η1,,ηn)(\eta_{1},\ldots,\eta_{n}) for (F,h)(F,h) coincides with normal coordinates at y0y_{0}, the spin connection on ΣF\Sigma F is ηiFν=dν(ηi)\nabla_{\eta_{i}}^{F}\nu=d\nu(\eta_{i}) (see (3.3)). Now ζi=efmηi\zeta_{i}=e^{\frac{f}{m}}\eta_{i}, and ηi,ν\eta_{i},\nu are identified with their images under π2\pi_{2}^{*}, so

(3.34) ζiZν=dν(ζi)\displaystyle\nabla^{Z}_{\zeta_{i}}\nu=d\nu(\zeta_{i}) =efmdν(ηi)=efmηiFν=ζiFν.\displaystyle=e^{\frac{f}{m}}d\nu(\eta_{i})=e^{\frac{f}{m}}\nabla^{F}_{\eta_{i}}\nu=\nabla^{F}_{\zeta_{i}}\nu.

Combining this with (3.33) gives (3.26). If nn is even, then by (3.25), (3.26) and (3.14) we have

(3.35) D¯ψ\displaystyle\bar{D}\psi =α=1nξα¯ξαψ+i=1mζi¯ζiψ\displaystyle=\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot\bar{\nabla}_{\xi_{\alpha}}\psi+\sum_{i=1}^{m}\zeta_{i}\cdot\bar{\nabla}_{\zeta_{i}}\psi
(3.36) =α=1n(ξαξαϕ)ν+i=1mζi(ϕζiν+12ζifψ)\displaystyle=\sum_{\alpha=1}^{n}(\xi_{\alpha}\cdot\nabla_{\xi_{\alpha}}\phi)\otimes\nu+\sum_{i=1}^{m}\zeta_{i}\cdot\left(\phi\otimes\nabla_{\zeta_{i}}\nu+\frac{1}{2}\zeta_{i}\cdot\nabla f\cdot\psi\right)
(3.37) =Dϕν+i=1mϕ¯(ζi^ζiν)12fψ\displaystyle=D\phi\otimes\nu+\sum_{i=1}^{m}\bar{\phi}\otimes(\zeta_{i}\hat{\,\cdot\,}\nabla_{\zeta_{i}}\nu)-\frac{1}{2}\nabla f\cdot\psi
(3.38) =Dϕν12(fϕ)ν+efmϕ¯i=1mηiηiFν\displaystyle=D\phi\otimes\nu-\frac{1}{2}(\nabla f\cdot\phi)\otimes\nu+e^{\frac{f}{m}}\bar{\phi}\otimes\sum_{i=1}^{m}\eta_{i}\cdot\nabla^{F}_{\eta_{i}}\nu
(3.39) =Dfϕν+efmϕ¯DFν,\displaystyle=D_{f}\phi\otimes\nu+e^{\frac{f}{m}}\bar{\phi}\otimes D^{F}\nu,

where the third equality also uses that Σ¯(M×F)\bar{\Sigma}(M\times F) is a module over l(T(M×F),g¯)\mathbb{C}\mathrm{l}(T(M\times F),\bar{g}) where ζi\zeta_{i} has unit length. This proves (3.27) in the case that nn is even. The other cases are proved similarly, by modifying the above computation according to the Clifford multiplication rules (3.14). ∎

4. Applications of spinors on the warped product, and more

Throughout this section, (Mn,g)(M^{n},g) and (Fm,h)(F^{m},h) are assumed to be spin manifolds, with FF being closed, scalar-flat, of unit volume, and admitting a nonzero parallel spinor ν\nu. For instance, we can take FF to be a flat torus with the appropriate spin structure. Using a weight f𝒞(M)f\in\mathcal{C}^{\infty}(M), form the warped product (M×F,g¯=ge2fmh)(M\times F,\bar{g}=g\oplus e^{-\frac{2f}{m}}h). Barred quantities denote those on the warped product. Spin geometry on the warped product was studied in §3, and we reuse the notations from there.

4.1. A spin proof of our positive mass theorem

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m), mm\in\mathbb{N} be an AE SMMS. We present an alternative proof of the following special case of Theorem 1.6 using spinors on the warped product.

Corollary 4.1.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) be an AE SMMS of order τ>n22\tau>\frac{n-2}{2} with mm\in\mathbb{N}, such that MM is spin. Also assume f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M) and RfL1(M,g)R_{f}\in L^{1}(M,g). If Rfm0R^{m}_{f}\geq 0, then 𝔪()0\mathfrak{m}(\mathcal{M})\geq 0, with equality if and only if (Mn,g)(M^{n},g) is isometric to (n,δij)(\mathbb{R}^{n},\delta_{ij}) and f0f\equiv 0.

Since (Mn,g)(M^{n},g) is AE with n>2n>2, MM_{\infty} is simply connected so the spin structure on (Mn,g)(M^{n},g) restricts to the trivial one on MM_{\infty}. This induces a trivialization of the spinor bundle ΣM\Sigma M over MM_{\infty}, and hence a (partial) trivialization of Σ¯(M×F)\bar{\Sigma}(M\times F) over M×FM_{\infty}\times F. A constant spinor means a spinor which is constant in these trivializations. The Hölder spaces 𝒞τk,α(Σ¯(M×F))\mathcal{C}^{k,\alpha}_{-\tau}(\bar{\Sigma}(M\times F)) are also defined in the obvious way, referring to the asymptotic behavior on MM.

We first construct Witten-type spinors in the spinor bundle Σ¯(M×F)\bar{\Sigma}(M\times F) of (M×F,g¯)(M\times F,\bar{g}).

Lemma 4.2.

Let (Mn,g,efdVolg,m)(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m), mm\in\mathbb{N} be an SMMS satisfying the hypotheses of Corollary 4.1. Then there exists a spinor ψΓ(Σ¯(M×F))\psi\in\Gamma(\bar{\Sigma}(M\times F)) such that

  • ψ=π1ϕπ2ν\psi=\pi_{1}^{*}\phi\otimes\pi_{2}^{*}\nu for some ϕΓ(𝚺M)\phi\in\Gamma({}^{\diamond}\mathbf{\Sigma}M) and νΓ(ΣF)\nu\in\Gamma(\Sigma F),

  • Dfϕ=0D_{f}\phi=0, and ν\nu is a unit norm parallel spinor,

  • D¯ψ=0\bar{D}\psi=0, and

  • There exists ψ0Γ(Σ¯(M×F))\psi_{0}\in\Gamma(\bar{\Sigma}(M\times F)) such that ψ0\psi_{0} is constant with unit norm on M×FM_{\infty}\times F, and ψψ0𝒞τ2,α(Σ¯(M×F))\psi-\psi_{0}\in\mathcal{C}^{2,\alpha}_{-\tau}(\bar{\Sigma}(M\times F)).

Proof.

The hypotheses allow Theorem 2.5 in [6] to be applied, giving a spinor ϕΓ(ΣM)\phi\in\Gamma(\Sigma M) such that

  • Dfϕ=0D_{f}\phi=0, where DfD_{f} is the weighted Dirac operator on MM, and

  • There exists ϕ0Γ(ΣM)\phi_{0}\in\Gamma(\Sigma M) such that ϕ0\phi_{0} is constant with unit norm on MM_{\infty}, and ϕϕ0𝒞τ2,α(ΣM)\phi-\phi_{0}\in\mathcal{C}^{2,\alpha}_{-\tau}(\Sigma M).

Take a parallel spinor νΓ(ΣF)\nu\in\Gamma(\Sigma F) with |ν|ΣF=1|\nu|_{\Sigma F}=1. Then |π2ν|ΣV=1|\pi_{2}^{*}\nu|_{\Sigma V}=1 by Lemma 3.1. For simplicity, assume nn and mm are not both odd. Define

(4.1) ψ=π1ϕπ2ν,ψ0=π1ϕ0π2ν,\psi=\pi_{1}^{*}\phi\otimes\pi_{2}^{*}\nu,\quad\psi_{0}=\pi_{1}^{*}\phi_{0}\otimes\pi_{2}^{*}\nu,

which are sections of π1(ΣM)ΣV=Σ¯(M×F)\pi_{1}^{*}(\Sigma M)\otimes\Sigma V=\bar{\Sigma}(M\times F) (see Proposition 3.2). Then ψ0\psi_{0} is constant on M×FM_{\infty}\times F, and Theorem 3.5 gives D¯ψ=0\bar{D}\psi=0. Moreover, by Proposition 3.3 we have |ψ0|Σ¯(M×F)=|ϕ0|ΣM|ν|ΣV=1|\psi_{0}|_{\bar{\Sigma}(M\times F)}=|\phi_{0}|_{\Sigma M}|\nu|_{\Sigma V}=1 on M×FM_{\infty}\times F. Similarly |ψψ0|Σ¯(M×F)=|ϕϕ0|ΣM|\psi-\psi_{0}|_{\bar{\Sigma}(M\times F)}=|\phi-\phi_{0}|_{\Sigma M}. Since ϕϕ0𝒞τ2,α(ΣM)\phi-\phi_{0}\in\mathcal{C}^{2,\alpha}_{-\tau}(\Sigma M), we have ψψ0𝒞τ2,α(Σ¯(M×F))\psi-\psi_{0}\in\mathcal{C}^{2,\alpha}_{-\tau}(\bar{\Sigma}(M\times F)), as desired.

In the remaining case where nn and mm are both odd, the argument is the same, except in (4.1) we replace ϕ\phi and ϕ0\phi_{0} by ϕ0\phi\oplus 0 and ϕ00\phi_{0}\oplus 0 respectively to accommodate the fact that 𝚺M=ΣMΣM{}^{\diamond}\mathbf{\Sigma}M=\Sigma M\oplus\Sigma M. ∎

Proof of Corollary 4.1.

Let =(Mn,g,efdVolg,m)\mathcal{M}=(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m), mm\in\mathbb{N} be an AE SMMS satisfying the hypotheses of the corollary. Form the warped product (M×F,g¯)(M\times F,\bar{g}), which has scalar curvature RfmR^{m}_{f}. Let ψ=π1ϕπ2νΓ(Σ¯(M×F))\psi=\pi_{1}^{*}\phi\otimes\pi_{2}^{*}\nu\in\Gamma(\bar{\Sigma}(M\times F)) be the Witten-type spinor provided by Lemma 4.2. The Lichnerowicz formula (3.5) and symmetry of D¯\bar{D} give

(4.2) 0=|D¯ψ|2=¯¯ψ,ψΣ¯(M×F)+14Rfm|ψ|2.0=|\bar{D}\psi|^{2}=\langle\bar{\nabla}^{*}\bar{\nabla}\psi,\psi\rangle_{\bar{\Sigma}(M\times F)}+\frac{1}{4}R^{m}_{f}|\psi|^{2}.

Write ξ=ψ0ψ𝒞τ2,α(Σ¯(M×F))\xi=\psi_{0}-\psi\in\mathcal{C}^{2,\alpha}_{-\tau}(\bar{\Sigma}(M\times F)). Integrating the above by parts over BρM×FB^{M}_{\rho}\times F, we get

(4.3) BρM×F(|¯ψ|2+14Rfm|ψ|2)𝑑Volg¯=ReSρM×Fψ,¯𝐧ψ𝑑Ag¯=Rei=1nSρM×Fψ,¯eiψei𝑑Volg¯\displaystyle\int_{B^{M}_{\rho}\times F}(|\bar{\nabla}\psi|^{2}+\frac{1}{4}R^{m}_{f}|\psi|^{2})d\mathrm{Vol}_{\bar{g}}=\mathrm{Re}\int_{S^{M}_{\rho}\times F}\langle\psi,\bar{\nabla}_{\mathbf{n}}\psi\rangle dA_{\bar{g}}=\mathrm{Re}\sum_{i=1}^{n}\int_{S^{M}_{\rho}\times F}\langle\psi,\bar{\nabla}_{e_{i}}\psi\rangle e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}}
(4.4) =Rei=1nSρM×F(ψ0,¯eiψ0ψ0,¯eiξξ,¯eiψ0+ξ,¯eiξ)ei𝑑Volg¯,\displaystyle\qquad=\mathrm{Re}\sum_{i=1}^{n}\int_{S^{M}_{\rho}\times F}\left(\langle\psi_{0},\bar{\nabla}_{e_{i}}\psi_{0}\rangle-\langle\psi_{0},\bar{\nabla}_{e_{i}}\xi\rangle-\langle\xi,\bar{\nabla}_{e_{i}}\psi_{0}\rangle+\langle\xi,\bar{\nabla}_{e_{i}}\xi\rangle\right)e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}},

where (ei)i=1n(e_{i})_{i=1}^{n} is a local orthonormal frame for TMTM. Since MM is AE and FF is closed, essentially the same arguments as in [25, Appendix A] show that the first, third and fourth terms on the right vanish as ρ\rho\to\infty. On the other hand, completing (ei)i=1n(e_{i})_{i=1}^{n} to a g¯\bar{g}-orthonormal frame for M×FM\times F, the second term is

(4.5) Rei=1nSρM×Fψ0,¯eiξei𝑑Volg¯\displaystyle\mathrm{Re}\sum_{i=1}^{n}\int_{S^{M}_{\rho}\times F}\langle\psi_{0},\bar{\nabla}_{e_{i}}\xi\rangle e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}} =14i=1na=1n+mSρM×F(ag¯aiig¯aa+𝒪(ρ2τ1))|ψ0|2ei𝑑Volg¯.\displaystyle=-\frac{1}{4}\sum_{i=1}^{n}\sum_{a=1}^{n+m}\int_{S^{M}_{\rho}\times F}(\partial_{a}\bar{g}_{ai}-\partial_{i}\bar{g}_{aa}+\mathcal{O}(\rho^{-2\tau-1}))|\psi_{0}|^{2}e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}}.

Since g¯ai0\bar{g}_{ai}\equiv 0 whenever ini\leq n and a>na>n, this becomes

(4.6) Rei=1nSρM×Fψ0,¯eiξei𝑑Volg¯\displaystyle\mathrm{Re}\sum_{i=1}^{n}\int_{S^{M}_{\rho}\times F}\langle\psi_{0},\bar{\nabla}_{e_{i}}\xi\rangle e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}} =14SρM×F(jg¯jiig¯aa+𝒪(ρ2τ1))|ψ0|2ei𝑑Volg¯\displaystyle=-\frac{1}{4}\int_{S^{M}_{\rho}\times F}(\partial_{j}\bar{g}_{ji}-\partial_{i}\bar{g}_{aa}+\mathcal{O}(\rho^{-2\tau-1}))|\psi_{0}|^{2}e_{i}\lrcorner d\mathrm{Vol}_{\bar{g}}

where the indices i,ji,j run over the frame for MM and the index aa runs over the full frame for M×FM\times F. Substituting this into (4.4), taking ρ\rho\to\infty, and using that |ψ0|=1|\psi_{0}|=1 outside a compact set, we get

(4.7) M×F(|¯ψ|2+14Rfm|ψ|2)𝑑Volg¯=limρ14SρM×F(ig¯ijjg¯aa)ej𝑑Volg¯=14𝔪(),\displaystyle\int_{M\times F}(|\bar{\nabla}\psi|^{2}+\frac{1}{4}R^{m}_{f}|\psi|^{2})d\mathrm{Vol}_{\bar{g}}=\lim_{\rho\to\infty}\frac{1}{4}\int_{S^{M}_{\rho}\times F}(\partial_{i}\bar{g}_{ij}-\partial_{j}\bar{g}_{aa})e_{j}\lrcorner d\mathrm{Vol}_{\bar{g}}=\frac{1}{4}\mathfrak{m}(\mathcal{M}),

where the last equality is by Definition 2.6. From this we see that 𝔪()0\mathfrak{m}(\mathcal{M})\geq 0 if Rfm0R^{m}_{f}\geq 0.

Working in a split orthonormal frame (ξ1,,ξn,ζ1,,ζm)(\xi_{1},\ldots,\xi_{n},\zeta_{1},\ldots,\zeta_{m}) centered at (x0,y0)M×F(x_{0},y_{0})\in M\times F, Theorem 3.5 gives that at (x0,y0)(x_{0},y_{0}),

(4.8) |¯ψ|2\displaystyle|\bar{\nabla}\psi|^{2} =α=1n|¯ξαψ|2+i=1m|¯ζiψ|2\displaystyle=\sum_{\alpha=1}^{n}|\bar{\nabla}_{\xi_{\alpha}}\psi|^{2}+\sum_{i=1}^{m}|\bar{\nabla}_{\zeta_{i}}\psi|^{2}
(4.9) =α=1n|ξαϕν|2+i=1m|ϕζiZν+12mζifψ|2.\displaystyle=\sum_{\alpha=1}^{n}|\nabla_{\xi_{\alpha}}\phi\otimes\nu|^{2}+\sum_{i=1}^{m}\left|\phi\otimes\nabla_{\zeta_{i}}^{Z}\nu+\frac{1}{2m}\zeta_{i}\cdot\nabla f\cdot\psi\right|^{2}.

Since ζiZν=efmζiFν\nabla_{\zeta_{i}}^{Z}\nu=e^{\frac{f}{m}}\nabla_{\zeta_{i}}^{F}\nu (see (3.34)), ν\nu is a unit norm parallel spinor, and ζi\zeta_{i} has unit norm with respect to g¯\bar{g}, it follows that

(4.10) |¯ψ|2\displaystyle|\bar{\nabla}\psi|^{2} =α=1n|ξαϕν|2+i=1m|12mζifψ|2\displaystyle=\sum_{\alpha=1}^{n}|\nabla_{\xi_{\alpha}}\phi\otimes\nu|^{2}+\sum_{i=1}^{m}\left|\frac{1}{2m}\zeta_{i}\cdot\nabla f\cdot\psi\right|^{2}
(4.11) =α=1n|ξαϕν|2+14m|f|2|ψ|2\displaystyle=\sum_{\alpha=1}^{n}|\nabla_{\xi_{\alpha}}\phi\otimes\nu|^{2}+\frac{1}{4m}|\nabla f|^{2}|\psi|^{2}
(4.12) =|ϕ|2+14m|f|2|ϕ|2.\displaystyle=|\nabla\phi|^{2}+\frac{1}{4m}|\nabla f|^{2}|\phi|^{2}.

The last equality is by Proposition 3.3, which also implies |ψ|=|ϕ||ν|=|ϕ||\psi|=|\phi||\nu|=|\phi|. Now suppose Rfm0R^{m}_{f}\geq 0 and 𝔪()=0\mathfrak{m}(\mathcal{M})=0. Then (4.7) implies |¯ψ|2=0|\bar{\nabla}\psi|^{2}=0, which in turn implies f=0\nabla f=0 and ϕ=0\nabla\phi=0 by (4.12). Thus ff is constant, but since f𝒞τ2,α(M)f\in\mathcal{C}^{2,\alpha}_{-\tau}(M), it is identically zero. Applying the Ricci identity (4.24) below to the DfD_{f}-harmonic spinor ϕ\phi implies (M,g)(M,g) is Ricci flat. By the Bishop–Gromov comparison theorem, any Ricci flat AE manifold is exactly Euclidean, so (Mn,g)(n,δij)(M^{n},g)\cong(\mathbb{R}^{n},\delta_{ij}). ∎

4.2. Weighted spin geometry identities

In the last subsection, we used Theorem 3.5 and an analog of Witten’s proof of the unweighted mass formula (1.3) applied to the warped product (M×F,g¯)(M\times F,\bar{g}), to get the formula (4.7) for the mass of \mathcal{M}. We can use this to give a new proof of the weighted Witten formula (1.5) of Baldauf and Ozuch. To see this, use (4.7), (4.12) and the fact that |ψ|=|ϕ||\psi|=|\phi| to get

(4.13) 14𝔪()\displaystyle\frac{1}{4}\mathfrak{m}(\mathcal{M}) =M×F(|ϕ|2+14m|f|2|ϕ|2+14Rfm|ϕ|2)𝑑Volg¯\displaystyle=\int_{M\times F}\left(|\nabla\phi|^{2}+\frac{1}{4m}|\nabla f|^{2}|\phi|^{2}+\frac{1}{4}R^{m}_{f}|\phi|^{2}\right)d\mathrm{Vol}_{\bar{g}}
(4.14) =M×F(|ϕ|2+14Rf|ϕ|2)𝑑Volg¯.\displaystyle=\int_{M\times F}\left(|\nabla\phi|^{2}+\frac{1}{4}R_{f}|\phi|^{2}\right)d\mathrm{Vol}_{\bar{g}}.

Since (Fm,h)(F^{m},h) has unit volume, dVolg¯=efdVolgdVolhd\mathrm{Vol}_{\bar{g}}=e^{-f}d\mathrm{Vol}_{g}\,d\mathrm{Vol}_{h}, and 𝔪()=𝔪f(g)\mathfrak{m}(\mathcal{M})=\mathfrak{m}_{f}(g), it follows that

(4.15) 𝔪f(g)=𝔪()=4M×F(|ϕ|2+14Rf|ϕ|2)ef𝑑Volg𝑑Volh=4M(|ϕ|2+14Rf|ϕ|2)ef𝑑Volg.\mathfrak{m}_{f}(g)=\mathfrak{m}(\mathcal{M})=4\int_{M\times F}\left(|\nabla\phi|^{2}+\frac{1}{4}R_{f}|\phi|^{2}\right)e^{-f}d\mathrm{Vol}_{g}\,d\mathrm{Vol}_{h}=4\int_{M}\left(|\nabla\phi|^{2}+\frac{1}{4}R_{f}|\phi|^{2}\right)e^{-f}d\mathrm{Vol}_{g}.

Recall that ϕΓ(ΣM)\phi\in\Gamma(\Sigma M) is a weighted Witten spinor, so the above formula is indeed (1.5).

The same outline allows us to reprove the weighted Lichnerowicz formula and the weighted Ricci identity. This is done in the next two propositions.

Proposition 4.3 (Weighted Lichnerowicz [6, 10]).

Let (Mn,g,f)(M^{n},g,f) be a weighted spin manifold. Then

(4.16) Df2ϕ=Δfϕ+14RfϕD_{f}^{2}\phi=-\Delta_{f}\phi+\frac{1}{4}R_{f}\phi

for all ϕΓ(ΣM)\phi\in\Gamma(\Sigma M), where Δf=f\Delta_{f}=-\nabla^{*}\nabla-\nabla_{\nabla f} is the weighted Laplacian acting on spinors.

Proof.

Form (M×F,g¯)(M\times F,\bar{g}) as before, and let νΓ(ΣF)\nu\in\Gamma(\Sigma F) be a nonzero parallel spinor. Applying the Lichnerowicz formula (3.5) to the spinor ψ=ϕνΓ(Σ¯(M×F))\psi=\phi\otimes\nu\in\Gamma(\bar{\Sigma}(M\times F)), and recalling that g¯\bar{g} has scalar curvature RfmR^{m}_{f}, we get

(4.17) D¯2ψ\displaystyle\bar{D}^{2}\psi =Δ¯ψ+14Rfmψ.\displaystyle=-\bar{\Delta}\psi+\frac{1}{4}R^{m}_{f}\psi.

Take a split orthonormal frame (ξ1,,ξn,ζ1,,ζm)(\xi_{1},\ldots,\xi_{n},\zeta_{1},\ldots,\zeta_{m}) for (T(M×F),g¯)(T(M\times F),\bar{g}) centered at (x0,y0)M×F(x_{0},y_{0})\in M\times F. Computing at (x0,y0)(x_{0},y_{0}), using the identities (3.22), (3.23) which hold there, and using Theorem 3.5, we have

(4.18) Δ¯ψ\displaystyle\bar{\Delta}\psi =α=1n(¯ξα¯ξαψ¯¯ξαξαψ)+i=1m(¯ζi¯ζiψ¯¯ζiζiψ)\displaystyle=\sum_{\alpha=1}^{n}(\bar{\nabla}_{\xi_{\alpha}}\bar{\nabla}_{\xi_{\alpha}}\psi-\bar{\nabla}_{\bar{\nabla}_{\xi_{\alpha}}\xi_{\alpha}}\psi)+\sum_{i=1}^{m}(\bar{\nabla}_{\zeta_{i}}\bar{\nabla}_{\zeta_{i}}\psi-\bar{\nabla}_{\bar{\nabla}_{\zeta_{i}}\zeta_{i}}\psi)
(4.19) =α=1nξαξαϕν+i=1m(14m2ζifζifψ1m¯fψ)\displaystyle=\sum_{\alpha=1}^{n}\nabla_{\xi_{\alpha}}\nabla_{\xi_{\alpha}}\phi\otimes\nu+\sum_{i=1}^{m}\left(\frac{1}{4m^{2}}\zeta_{i}\cdot\nabla f\cdot\zeta_{i}\cdot\nabla f\cdot\psi-\frac{1}{m}\bar{\nabla}_{\nabla f}\psi\right)
(4.20) =Δϕν14m|f|2ψfϕν\displaystyle=\Delta\phi\otimes\nu-\frac{1}{4m}|\nabla f|^{2}\psi-\nabla_{\nabla f}\phi\otimes\nu
(4.21) =Δfϕν14m|f|2ψ.\displaystyle=\Delta_{f}\phi\otimes\nu-\frac{1}{4m}|\nabla f|^{2}\psi.

On the other hand, Theorem 3.5 also gives D¯2ψ=Df2ϕν\bar{D}^{2}\psi=D_{f}^{2}\phi\otimes\nu. Thus (4.17) becomes

(4.22) (Df2ϕν)\displaystyle(D_{f}^{2}\phi\otimes\nu) =Δfϕν+14m|f|2ψ+14(Rf1m|f|2)ψ\displaystyle=-\Delta_{f}\phi\otimes\nu+\frac{1}{4m}|\nabla f|^{2}\psi+\frac{1}{4}\left(R_{f}-\frac{1}{m}|\nabla f|^{2}\right)\psi
(4.23) =(Δfϕ+14Rfϕ)ν.\displaystyle=\left(-\Delta_{f}\phi+\frac{1}{4}R_{f}\phi\right)\otimes\nu.

The proposition follows from this since ν0\nu\neq 0 is parallel and hence nonvanishing. ∎

Proposition 4.4 (Weighted Ricci [6]).

Let (Mn,g,f)(M^{n},g,f) be a weighted spin manifold. Then

(4.24) [Df,X]ϕ\displaystyle[D_{f},\nabla_{X}]\phi =12Ricf(X)ϕ\displaystyle=\frac{1}{2}\mathrm{Ric}_{f}(X)\cdot\phi

for all XTMX\in TM and ϕΓ(ΣM)\phi\in\Gamma(\Sigma M).

Proof.

Reuse the setup in the proof of Proposition 4.3. The usual Ricci identity (see e.g. [9, Corollary 2.8]) applied on Σ¯(M×F)\bar{\Sigma}(M\times F) yields

(4.25) 12Ricg¯(X)\displaystyle-\frac{1}{2}\mathrm{Ric}_{\bar{g}}(X)\cdot =α=1nξα¯X,ξα+i=1mζi¯X,ζi,\displaystyle=\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot\bar{\mathcal{R}}_{X,\xi_{\alpha}}+\sum_{i=1}^{m}\zeta_{i}\cdot\bar{\mathcal{R}}_{X,\zeta_{i}},

where ¯Y,Z:=¯Y¯Z¯Z¯Y¯[Y,Z]\bar{\mathcal{R}}_{Y,Z}:=\bar{\nabla}_{Y}\bar{\nabla}_{Z}-\bar{\nabla}_{Z}\bar{\nabla}_{Y}-\bar{\nabla}_{[Y,Z]}. We claim that Ricg¯(X)=Ricfm(X)\mathrm{Ric}_{\bar{g}}(X)=\mathrm{Ric}^{m}_{f}(X). Indeed, RicM×F(X)\mathrm{Ric}_{M\times F}(X) is a horizontal vector because RicM×F(X,Y)=0\mathrm{Ric}_{M\times F}(X,Y)=0 for all vertical vectors; see e.g. [8, Proposition 9.106].Using that g=g¯g=\bar{g} and Ricg¯=Ricfm\mathrm{Ric}_{\bar{g}}=\mathrm{Ric}^{m}_{f} on horizontal vectors, as well as the definitions of Ricg¯,Ricfm\mathrm{Ric}_{\bar{g}},\mathrm{Ric}^{m}_{f}, we see that

(4.26) g(Ricg¯(X),X)=g¯(Ricg¯(X),X)=Ricg¯(X,X)=Ricfm(X,X)=g(Ricfm(X),X)g(\mathrm{Ric}_{\bar{g}}(X),X^{\prime})=\bar{g}(\mathrm{Ric}_{\bar{g}}(X),X^{\prime})=\mathrm{Ric}_{\bar{g}}(X,X^{\prime})=\mathrm{Ric}^{m}_{f}(X,X^{\prime})=g(\mathrm{Ric}^{m}_{f}(X),X^{\prime})

for all horizontal vectors XTMX^{\prime}\in TM. Thus Ricg¯(X)=Ricfm(X)\mathrm{Ric}_{\bar{g}}(X)=\mathrm{Ric}^{m}_{f}(X).

Since (ξ1,,ξn)(\xi_{1},\ldots,\xi_{n}) coincide with normal coordinate vector fields for MM at x0x_{0}, we compute at that point

(4.27) α=1nξα¯X,ξα+i=1mζi¯X,ζi\displaystyle\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot\bar{\mathcal{R}}_{X,\xi_{\alpha}}+\sum_{i=1}^{m}\zeta_{i}\cdot\bar{\mathcal{R}}_{X,\zeta_{i}} =α=1nξα(¯X¯ξα¯ξα¯X)+i=1mζi(¯X¯ζi¯ζi¯X¯[X,ζi])\displaystyle=\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot(\bar{\nabla}_{X}\bar{\nabla}_{\xi_{\alpha}}-\bar{\nabla}_{\xi_{\alpha}}\bar{\nabla}_{X})+\sum_{i=1}^{m}\zeta_{i}\cdot(\bar{\nabla}_{X}\bar{\nabla}_{\zeta_{i}}-\bar{\nabla}_{\zeta_{i}}\bar{\nabla}_{X}-\bar{\nabla}_{[X,\zeta_{i}]})
(4.28) =α=1n(¯X(ξα¯ξα)ξα¯ξα¯X¯Xξα¯ξα)\displaystyle=\sum_{\alpha=1}^{n}(\bar{\nabla}_{X}(\xi_{\alpha}\cdot\bar{\nabla}_{\xi_{\alpha}})-\xi_{\alpha}\cdot\bar{\nabla}_{\xi_{\alpha}}\bar{\nabla}_{X}-\bar{\nabla}_{X}\xi_{\alpha}\cdot\bar{\nabla}_{\xi_{\alpha}})
(4.29) +i=1m(¯X(ζi¯ζi)ζi¯ζi¯X¯Xζi¯ζiζi¯[X,ζi]).\displaystyle\quad+\sum_{i=1}^{m}(\bar{\nabla}_{X}(\zeta_{i}\cdot\bar{\nabla}_{\zeta_{i}})-\zeta_{i}\cdot\bar{\nabla}_{\zeta_{i}}\bar{\nabla}_{X}-\bar{\nabla}_{X}\zeta_{i}\cdot\bar{\nabla}_{\zeta_{i}}-\zeta_{i}\cdot\bar{\nabla}_{[X,\zeta_{i}]}).

By (3.22), we have ¯Xξα=¯Xζi=0\bar{\nabla}_{X}\xi_{\alpha}=\bar{\nabla}_{X}\zeta_{i}=0 and [X,ζi]=¯Xζi¯ζiX=1m(Xf)ζi[X,\zeta_{i}]=\bar{\nabla}_{X}\zeta_{i}-\bar{\nabla}_{\zeta_{i}}X=\frac{1}{m}(Xf)\zeta_{i}. Thus

(4.30) α=1nξα¯X,ξα+i=1mζi¯X,ζi\displaystyle\sum_{\alpha=1}^{n}\xi_{\alpha}\cdot\bar{\mathcal{R}}_{X,\xi_{\alpha}}+\sum_{i=1}^{m}\zeta_{i}\cdot\bar{\mathcal{R}}_{X,\zeta_{i}} =¯XD¯D¯¯X1m(Xf)i=1mζi¯ζi.\displaystyle=\bar{\nabla}_{X}\bar{D}-\bar{D}\bar{\nabla}_{X}-\frac{1}{m}(Xf)\sum_{i=1}^{m}\zeta_{i}\cdot\bar{\nabla}_{\zeta_{i}}.

Applying this to ψ=ϕν\psi=\phi\otimes\nu on both sides, then using (4.25) on the left and Theorem 3.5 on the right, we get

(4.31) 12Ricg¯(X)ψ\displaystyle-\frac{1}{2}\mathrm{Ric}_{\bar{g}}(X)\cdot\psi =[Df,X]ϕν1m(Xf)i=1mζi(12mζifψ)\displaystyle=-[D_{f},\nabla_{X}]\phi\otimes\nu-\frac{1}{m}(Xf)\sum_{i=1}^{m}\zeta_{i}\cdot\left(\frac{1}{2m}\zeta_{i}\cdot\nabla f\cdot\psi\right)
(4.32) =[Df,X]ϕν+12m(Xf)fψ.\displaystyle=-[D_{f},\nabla_{X}]\phi\otimes\nu+\frac{1}{2m}(Xf)\nabla f\cdot\psi.

Since Ricg¯(X)=Ricfm(X)\mathrm{Ric}_{\bar{g}}(X)=\mathrm{Ric}^{m}_{f}(X), this gives

(4.33) [Df,X]ϕν\displaystyle[D_{f},\nabla_{X}]\phi\otimes\nu =[12(Ricfm(X)1m(dfdf)(X))ϕ]ν=(12Ricf(X)ϕ)ν\displaystyle=\left[\frac{1}{2}\left(\mathrm{Ric}^{m}_{f}(X)-\frac{1}{m}(df\otimes df)(X)\right)\cdot\phi\right]\otimes\nu=\left(\frac{1}{2}\mathrm{Ric}_{f}(X)\cdot\phi\right)\otimes\nu

and the proposition follows. ∎

4.3. Dirac spectra of closed manifolds

Now assume additionally that (Mn,g)(M^{n},g) is compact without boundary. We will see what our earlier results imply for the spectrum of the Dirac operator on MM. Relationships between the Dirac spectra of MM and of fibrations over MM were previously studied in [2, 28, 32] (and references therein), although their settings are rather different from ours.

An SMMS (Mn,g,efdVolg,m)(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) is said to be quasi-Einstein if Ricfm=λg\mathrm{Ric}^{m}_{f}=\lambda g for some λ\lambda\in\mathbb{R} [11]. Our first result here characterizes this property in terms of the warped product Dirac operator D¯\bar{D}. Denote by λ1(D¯)\lambda_{1}(\bar{D}) the least eigenvalue of D¯\bar{D} in absolute value.

Proposition 4.5.

We have λ1(D¯)2n+m4(n+m1)minRfm\lambda_{1}(\bar{D})^{2}\geq\frac{n+m}{4(n+m-1)}\min R^{m}_{f} for all mm\in\mathbb{N}. Equality holds for mm\in\mathbb{N} if and only if the SMMS (Mn,g,efdVolg,m)(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) is quasi-Einstein.

Proof.

Since g¯\bar{g} has scalar curvature RfmR^{m}_{f}, Friedrich’s inequality [19] (also [9, Theorem 5.3]) applied to the warped product says that λ1(D¯)2n+m4(n+m1)minRfm\lambda_{1}(\bar{D})^{2}\geq\frac{n+m}{4(n+m-1)}\min R^{m}_{f}, with equality if and only if (M×F,g¯)(M\times F,\bar{g}) is Einstein. The latter holds if and only if (Mn,g,efdVolg,m)(M^{n},g,e^{-f}d\mathrm{Vol}_{g},m) is quasi-Einstein, since for horizontal vectors X,XTMX,X^{\prime}\in TM we have

(4.34) Ricg¯(X,X)=Ricfm(X,X).\mathrm{Ric}_{\bar{g}}(X,X^{\prime})=\mathrm{Ric}^{m}_{f}(X,X^{\prime}).

If Dϕ=λϕD\phi=\lambda\phi for some ϕΓ(ΣM)\phi\in\Gamma(\Sigma M), then D¯(ϕν)=λϕν\bar{D}(\phi\otimes\nu)=\lambda\phi\otimes\nu by Theorem 3.5, where νΓ(ΣF)\nu\in\Gamma(\Sigma F) is a nontrivial parallel spinor. Thus the spectrum of DD is contained in the spectrum of D¯\bar{D}, and Proposition 4.5 implies λ1(D)2n+m4(n+m1)minRfm\lambda_{1}(D)^{2}\geq\frac{n+m}{4(n+m-1)}\min R^{m}_{f} for all mm\in\mathbb{N}. This bound is similar to, but weaker than the weighted Friedrich inequality λ1(D)2n4(n1)minRf\lambda_{1}(D)^{2}\geq\frac{n}{4(n-1)}\min R_{f} [6, Theorem 1.23]. Equality in the latter inequality holds if and only if ff is constant, and (Mn,g)(M^{n},g) admits a nontrivial Killing spinor and is Einstein. The next corollary uses this to give a similar characterization of equality in our bound.

Corollary 4.6.

We have λ1(D)2n+m4(n+m1)minRfm\lambda_{1}(D)^{2}\geq\frac{n+m}{4(n+m-1)}\min R^{m}_{f} for all mm\in\mathbb{N} and f𝒞(M)f\in\mathcal{C}^{\infty}(M). Equality holds for some mm if and only if ff is constant and MM admits a nontrivial parallel spinor (hence is Ricci flat).

Proof.

Argue as in the last paragraph to get the claimed bound on λ1(D)2\lambda_{1}(D)^{2}. Now suppose equality holds in the bound for some mm\in\mathbb{N}. By the weighted Friedrich inequality, we have

(4.35) λ1(D)2n4(n1)minRfn+m4(n+m1)minRfm,\lambda_{1}(D)^{2}\geq\frac{n}{4(n-1)}\min R_{f}\geq\frac{n+m}{4(n+m-1)}\min R^{m}_{f},

but the equality assumption turns both inequalities above into equalities. The first equality implies ff is constant and MM is Einstein, so Rfm=Rf=RR^{m}_{f}=R_{f}=R. This, the second equality, and the fact that n4(n1)>n+m4(n+m1)\frac{n}{4(n-1)}>\frac{n+m}{4(n+m-1)} imply that minRf0\min R_{f}\leq 0. The first equality in turn forces λ1(D)=minRf=0\lambda_{1}(D)=\min R_{f}=0, so MM admits a nontrivial parallel spinor. Finally, a spin manifold admitting a nontrivial parallel spinor must be Ricci-flat (see e.g. [20, §3.2]).

Conversely, if ff is constant and MM admits a nontrivial parallel spinor, then λ1(D)2=0\lambda_{1}(D)^{2}=0. On the other hand (Mn,g)(M^{n},g) is Ricci-flat, hence Rfm=R=0R^{m}_{f}=R=0. Thus λ1(D)2=n+m4(n+m1)minRfm\lambda_{1}(D)^{2}=\frac{n+m}{4(n+m-1)}\min R^{m}_{f}. ∎

Specializing our discussion to harmonic spinors, first note that the weighted Lichnerowicz formula Df2=+14RfD_{f}^{2}=\nabla^{*}\nabla+\frac{1}{4}R_{f} implies the following, since RfRfmR_{f}\geq R^{m}_{f} for m>0m>0:

Corollary 4.7.

If there exist m>0m>0 and f𝒞(M)f\in\mathcal{C}^{\infty}(M) with Rfm0R^{m}_{f}\geq 0 and Rfm>0R^{m}_{f}>0 at some point, then MM admits no nontrivial harmonic spinors.

Corollary 1.8 extends Corollary 4.7 to include m<1nm<1-n; this is not implied by the weighted Lichnerowicz formula since RfmRfR^{m}_{f}\leq R_{f} here. For its proof, we make the following observation. In addition to (M,g)(M,g) and its spinor bundle ΣM\Sigma M, consider the conformally changed manifold (M,g~=e2fn1g)(M,\tilde{g}=e^{-\frac{2f}{n-1}}g) and its spinor bundle Σ~M\tilde{\Sigma}M. According to [9, §2.3.5], there is a natural bundle isometry Φ:ΣMΣ~M\Phi:\Sigma M\to\tilde{\Sigma}M, and the Dirac operators D,D~D,\tilde{D} of the respective bundles are related by

(4.36) Df=D12f=efn1Φ1D~Φ.D_{f}=D-\frac{1}{2}\nabla f\cdot=e^{-\frac{f}{n-1}}\Phi^{-1}\tilde{D}\Phi.

Thus, the weighted Dirac operator on (M,g,f)(M,g,f) is obtained from the usual Dirac operator on (M,g~)(M,\tilde{g}) by conjugating by Φ\Phi and multiplying by a function. Interestingly, g~\tilde{g} is the same metric used in §2 to prove our positive mass theorems.

Proof of Corollary 1.8.

By Lemma 2.4 (or (2.35)), the assumption that Rfm0R_{f}^{m}\geq 0 and is positive somewhere imply that the scalar curvature R~\tilde{R} of g~\tilde{g} is 0\geq 0 and is positive somewhere. The Lichnerowicz formula applied to the spinor bundle Σ~M\tilde{\Sigma}M of (M,g~)(M,\tilde{g}) shows that the Dirac operator D~\tilde{D} on Σ~M\tilde{\Sigma}M has trivial kernel. Then (4.36) shows that DfD_{f} has trivial kernel on ΣM\Sigma M. As observed independently in [6] and [10], DD and DfD_{f} have the same eigenvalues. Thus DD also has trivial kernel. ∎

Remark.

Corollary 1.8 cannot be extended to any m[1,0)m\in[-1,0). In fact, every closed manifold MM of dimension 3\geq 3 admits a metric gg and a function ff for which Rfm>0R^{m}_{f}>0 whenever m[1,0)m\in[-1,0). Indeed, by [17, Theorem 3] there exists a metric gg on MM with MR𝑑Volg>0\int_{M}R\,d\mathrm{Vol}_{g}>0, so the function

(4.37) u:=R2MR𝑑Volg2Vol(M,g)u:=\frac{R}{2}-\frac{\int_{M}R\,d\mathrm{Vol}_{g}}{2\mathrm{Vol}(M,g)}

satisfies Mu𝑑Volg=0\int_{M}u\,d\mathrm{Vol}_{g}=0. Thus u=Δfu=-\Delta f for some f𝒞(M)f\in\mathcal{C}^{\infty}(M), so for all m[1,0)m\in[-1,0) we have

(4.38) Rfm=R+2Δfm+1m|f|2R2u=MR𝑑VolgVol(M,g)>0.R^{m}_{f}=R+2\Delta f-\frac{m+1}{m}|\nabla f|^{2}\geq R-2u=\frac{\int_{M}R\,d\mathrm{Vol}_{g}}{\mathrm{Vol}(M,g)}>0.

Nonetheless it remains unclear whether Corollary 1.8 can be extended to also include all values of mm in the interval (1n,1)(1-n,-1).

Corollary 1.8 motivates the problem of finding f𝒞(M)f\in\mathcal{C}^{\infty}(M) and m[1n,0]m\in\mathbb{R}\setminus[1-n,0] such that RfmR^{m}_{f} is a constant μm\mu_{m}, since if μm>0\mu_{m}>0 then the corollary implies MM has no nontrivial harmonic spinors. The next proposition shows that this problem can be solved, and μm>0\mu_{m}>0 when R>0R>0. While no new obstructions to harmonic spinors arise from this, we will see shortly that the constants μm\mu_{m} relate to well-known lower bounds for the Dirac spectrum.

Proposition 4.8.

Let (M,g)(M,g) be a closed Riemannian manifold. For each m{0}m\in\mathbb{R}\setminus\{0\}, there is a unique constant μm\mu_{m}\in\mathbb{R} and a smooth function f𝒞(M)f\in\mathcal{C}^{\infty}(M) unique up to translation, such that Rfm=μmR^{m}_{f}=\mu_{m}. In fact

(4.39) μm=infvH1(M,g),vL2(M,g)=1M(4mm+1|v|2+Rv2)𝑑Volg.\mu_{m}=\inf_{\begin{subarray}{c}v\in H^{1}(M,g),\\ \lVert v\rVert_{L^{2}(M,g)}=1\end{subarray}}\int_{M}\left(\frac{4m}{m+1}|\nabla v|^{2}+Rv^{2}\right)\,d\mathrm{Vol}_{g}.
Proof.

Inspired by [16, Theorem 2], we set u:=em+12mfu:=e^{-\frac{m+1}{2m}f}. A careful computation yields

(4.40) 4mm+1Δgu+Ru=Rfmu.-\frac{4m}{m+1}\Delta_{g}u+Ru=R_{f}^{m}u.

Thus, solving Rfm=μmR^{m}_{f}=\mu_{m} is equivalent to finding a positive solution uu to the eigenvalue problem

(4.41) Lu:=(4mm+1Δg+R)u=μmu.Lu:=\left(-\frac{4m}{m+1}\Delta_{g}+R\right)u=\mu_{m}u.

By standard elliptic theory, only the principal eigenfunctions of LL do not change sign; thus, μm\mu_{m} must equal the principal eigenvalue, which is given by (4.39). Also the μm\mu_{m}-eigenspace is one-dimensional and consists of smooth functions. Thus, uu exists uniquely up to positive scaling, and ff is unique up to translation. ∎

Now let (Mn,g)(M^{n},g) be a closed spin manifold with Dirac operator DD. Write λ1()\lambda_{1}(\cdot) for the lowest eigenvalue of a second-order operator. Friedrich’s [19] and Hijazi’s [22] inequalities respectively imply that

(4.42) λ1(D)2\displaystyle\lambda_{1}(D)^{2} n4(n1)λ1(4Δ+R),\displaystyle\geq\frac{n}{4(n-1)}\lambda_{1}(-4\Delta+R),
(4.43) λ1(D)2\displaystyle\lambda_{1}(D)^{2} n4(n1)λ1(),\displaystyle\geq\frac{n}{4(n-1)}\lambda_{1}(\Box),

where =4(n1)n2Δ+R\Box=-\frac{4(n-1)}{n-2}\Delta+R is the conformal Laplacian. Meanwhile, observe using (4.39) that

(4.44) μmλ1(4Δ+R)\displaystyle\mu_{m}\uparrow\lambda_{1}(-4\Delta+R) as 0<m,\displaystyle\quad\text{as }0<m\uparrow\infty,
(4.45) μmλ1()\displaystyle\mu_{m}\uparrow\lambda_{1}(\Box) as <m1n.\displaystyle\quad\text{as }-\infty<m\uparrow 1-n.

Therefore, the inequalities (4.42) and (4.43) are implied by the family of weaker inequalities

(4.46) λ1(D)2n4(n1)μm,m[1n,0].\lambda_{1}(D)^{2}\geq\frac{n}{4(n-1)}\mu_{m},\quad m\in\mathbb{R}\setminus[1-n,0].

This family of inequalities interpolates between the inequalities of Friedrich and Hijazi.

References

  • [1] F. Abedin and J. Corvino. On the P-scalar curvature. J. Geom. Anal., 27(2):1589–1623, 2017.
  • [2] B. Ammann and C. Bär. The Dirac operator on nilmanifolds and collapsing circle bundles. Ann. Global Anal. Geom., 16:221–253, 1998.
  • [3] R. Arnowitt, S. Deser, and C. W. Misner. Canonical variables for general relativity. Phys. Rev., 117(6):1595, 1960.
  • [4] R. Arnowitt, S. Deser, and C. W. Misner. Energy and the criteria for radiation in general relativity. Phys. Rev., 118(4):1100, 1960.
  • [5] R. Arnowitt, S. Deser, and C. W. Misner. Coordinate invariance and energy expressions in general relativity. Phys. Rev., 122(3):997, 1961.
  • [6] J. Baldauf and T. Ozuch. Spinors and mass on weighted manifolds. Comm. Math. Phys., 394(3):1153–1172, 2022.
  • [7] R. Bartnik. The mass of an asymptotically flat manifold. Comm. Pure Appl. Math., 39(5):661–693, 1986.
  • [8] A. L. Besse. Einstein manifolds. Classics in Mathematics. Springer Berlin, Heidelberg, 1987.
  • [9] J.-P. Bourguignon, O. Hijazi, J.-L. Milhorat, A. Moroianu, and S. Moroianu. A spinorial approach to Riemannian and conformal geometry. EMS Monographs in Mathematics. European Mathematical Society, 2015.
  • [10] V. Branding and G. Habib. Eigenvalue estimates on weighted manifolds. Preprint, arXiv:2201.06375, 2022.
  • [11] J. S. Case. Smooth metric measure spaces and quasi-Einstein metrics. Internat. J. Math., 23(10):1250110, 2012.
  • [12] J. Chu and J. Zhu. A non-spin method to the positive weighted mass theorem for weighted manifolds. Preprint, arXiv:2305.12909, 2023.
  • [13] X. Dai. A positive mass theorem for spaces with asymptotic SUSY compactification. Comm. Math. Phys., 244(2):335–345, 2004.
  • [14] G. B. De Luca and A. Tomasiello. Leaps and bounds towards scale separation. J. High Energy Phys., 2021(86), 2021.
  • [15] J. Deng. Curvature-dimension condition meets Gromov’s nn-volumic scalar curvature. SIGMA Symmetry Integrability Geom. Methods Appl., 17:013, 2021.
  • [16] F. Dobarro and E. Lami Dozo. Scalar curvature and warped products of Riemann manifolds. Trans. Amer. Math. Soc., 303(1):161–168, 1987.
  • [17] H. I. Eliasson. On variations of metrics. Math. Scand., 29(2):317–327, 1971.
  • [18] E. M. Fan. Topology of three-manifolds with positive P-scalar curvature. Proc. Amer. Math. Soc., 136(9):3255–3261, 2008.
  • [19] T. Friedrich. Der erste Eigenwert des Dirac‐Operators einer kompakten, Riemannschen Mannigfaltigkeit nichtnegativer Skalarkrümmung. Math. Nachr., 97(1):117–146, 1980.
  • [20] T. Friedrich. Dirac operators in Riemannian geometry, volume 25 of Graduate Studies in Mathematics. American Mathematical Society, 2000.
  • [21] G. J. Galloway and E. Woolgar. Cosmological singularities in Bakry–Émery spacetimes. J. Geom. Phys., 86:359–369, 2014.
  • [22] O. Hijazi. A conformal lower bound for the smallest eigenvalue of the Dirac operator and Killing spinors. Comm. Math. Phys., 104(1):151–162, 1986.
  • [23] H. B. Lawson and M.-L. Michelsohn. Spin geometry, volume 38 of Princeton Mathematical Series. Princeton University Press, 1989.
  • [24] D. A. Lee. Geometric relativity, volume 201 of Graduate Studies in Mathematics. American Mathematical Society, 2019.
  • [25] J. M. Lee and T. H. Parker. The Yamabe problem. Bull. Amer. Math. Soc., 17(1):37–91, 1987.
  • [26] A. Lichnerowicz. Variétés riemanniennes à tenseur C non négatif. C. R. Acad. Sci. Paris Sér., 271:A650–A653, 1970.
  • [27] A. Lichnerowicz. Variétés kählériennes à première classe de Chern non negative et variétés riemanniennes à courbure de Ricci généralisée non negative. J. Differential Geom., 6(1):47–94, 1971.
  • [28] J. Lott. Collapsing and Dirac-type operators. Geom. Dedicata, 91:175–196, 2002.
  • [29] B. O’Neill. The fundamental equations of a submersion. Michigan Math. J., 13(4):459–469, 1966.
  • [30] T. Parker and C. H. Taubes. On Witten’s proof of the positive energy theorem. Comm. Math. Phys., 84(2):223–238, 1982.
  • [31] G. Perelman. The entropy formula for the Ricci flow and its geometric applications. Preprint, arXiv:math/0211159, 2002.
  • [32] S. Roos. The Dirac operator under collapse to a smooth limit space. Ann. Global Anal. Geom., 57(1):121–151, 2020.
  • [33] R. Schoen and S.-T. Yau. On the proof of the positive mass conjecture in general relativity. Comm. Math. Phys., 65(1):45–76, 1979.
  • [34] R. Schoen and S.-T. Yau. Proof of the positive mass theorem. II. Comm. Math. Phys., 79(2):231–260, 1979.
  • [35] M. Y. Wang. Parallel spinors and parallel forms. Ann. Global Anal. Geom., 7(1):59–68, 1989.
  • [36] E. Witten. A new proof of the positive energy theorem. Comm. Math. Phys., 80(3):381–402, 1981.
  • [37] E. Woolgar. Scalar–tensor gravitation and the Bakry–Émery–Ricci tensor. Classical Quantum Gravity, 30(8):085007, 2013.
  • [38] E. Woolgar and W. Wylie. Cosmological singularity theorems and splitting theorems for N-Bakry–Émery spacetimes. J. Math. Phys., 57(2):022504, 2016.
  • [39] P. Wu. A Weitzenböck formula for canonical metrics on four-manifolds. Trans. Amer. Math. Soc., 369(2):1079–1096, 2017.