Skip to main content
Log in

On the P-Scalar Curvature

  • Published:
The Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

We study the P-scalar curvature operator used by Perelman in the context of manifolds with density. We verify that the Gromov–Lawson surgery for positive scalar curvature extends naturally to positive P-scalar curvature. By studying the first variation of the P-scalar curvature operator, we obtain local perturbation results for the P-scalar curvature. As a corollary, we construct small perturbations (in the category of manifolds with density) of the round unit hemisphere \(\mathbb {S}^n_+\) which are compactly supported away from the boundary and yield manifolds with density with non-constant P-scalar curvature at least \(n(n-1)\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Akutagawa, K., Ishida, M., LeBrun, C.: Perelman’s invariant, Ricci flow, and the Yamabe invariants of smooth manifolds. Arch. Math. (Basel) 88(1), 71–76 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  2. Brendle, S., Marques, F.: Scalar curvature rigidity of geodesic balls in \(S^n\). J. Differ. Geom. 88(3), 379–394 (2011)

    MathSciNet  MATH  Google Scholar 

  3. Brendle, S., Marques, F., Neves, A.: Deformations of the hemisphere that increase scalar curvature. Invent. Math. 185(1), 175–197 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  4. Chang, A., Gursky, M., Yang, P.: Conformal invariants associated to a measure. Proc. Natl Acad. Sci. USA 103(8), 2535–2540 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chruściel, P.T., Delay, E.: On mapping properties of the general relativistic constraints operator in weighted function spaces, with applications. Mém. Soc. Math. Fr. (N.S.) No. 94 (2003)

  6. Corvino, J.: Scalar curvature deformation and a gluing construction for the Einstein constraint equations. Commun. Math. Phys. 214(1), 137–189 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  7. Corvino, J., Schoen, R.M.: On the asymptotics of the vacuum Einstein constraint equations. J. Differ. Geom. 73(2), 185–217 (2006)

    MathSciNet  MATH  Google Scholar 

  8. Corvino, J., Eichmair, M., Miao, P.: Deformation of scalar curvature and volume. Math. Ann. 357(2), 551–584 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  9. Fan, E.M.: Topology of three-manifolds with positive \(P\)-scalar curvature. Proc. Am. Math. Soc. 136(9), 3255–3261 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  10. Fang, F., Zhang, Y.: Perelman’s \(\lambda \)-functional and Seiberg–Witten equations. Front. Math. China 2(2), 191–210 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  11. Fischer, A.E., Marsden, J.E.: Deformations of the scalar curvature. Duke Math. J. 42(3), 519–547 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  12. Gromov, M., Lawson Jr., H.B.: The classification of simply connected manifolds of positive scalar curvature. Ann. Math. (2) 111(3), 423–434 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  13. Gromov, M., Lawson Jr., H.B.: Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math. 58, 83–196 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  14. Hatcher, A.: Notes on Basic 3-Manifold Topology. http://www.math.cornell.edu/hatcher

  15. Ho, P.T.: The structure of \(\phi \)-stable minimal hypersurfaces in manifolds of nonnegative \(P\)-scalar curvature. Math. Ann. 348(2), 319–332 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  16. Kahn, J., Markovic, V.: Immersing almost geodesic surfaces in a closed hyperbolic three-manifold. Ann. Math. (2) 175(3), 1127–1190 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  17. Lawson Jr., H.B., Michelson, M.-L.: Spin Geometry, Princeton Mathematical Series, vol. 38. Princeton University Press, Princeton (1989)

    Google Scholar 

  18. Ma, L., Zhu, A.: Eigenvalues and lambda constants on Riemannian submersions. Geom. Dedicata 129, 73–82 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  19. Mahmoudi, F., Mazzeo, R., Pacard, F.: Constant mean curvature hypersurfaces condensing on a submanifold. Geom. Funct. Anal. 16, 924–958 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  20. Perelman, G.: The entropy formula for the Ricci flow and its geometric applications. http://arXiv.org/abs/math.DG/0211159

  21. Schoen, R.M., Yau, S.-T.: Existence of incompressible minimal surfaces and the topology of three-manifolds with nonnegative scalar curvature. Ann. Math. (2) 110(1), 127–142 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  22. Schoen, R.M., Yau, S.-T.: On the structure of manifolds with positive scalar curvature. Manuscr. Math. 28(1–3), 159–183 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  23. Schoen, R.M., Yau, S.-T.: The structure of manifolds with positive scalar curvature. In: Directions in partial differential equations (Madison, WI, 1985), vol. 54, pp. 235–242. Publ. Math. Res. Center Univ, Wisconsin, Academic Press, Boston (1987)

  24. Wei, G., Wylie, W.: Comparison geometry for the Bakry–Emery Ricci tensor. J. Differ. Geom. 83(2), 377–405 (2009)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge partial support from NSF grants DMS-0707317 and DMS-1207844, as well as from the Lafayette College EXCEL Scholars program.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Justin Corvino.

Appendix

Appendix

1.1 Product Formulas for the Gradient and Laplacian

Let D be the open disk \(B_{r_0}(p)\) or the open tube \(\mathcal {N}_{r_0}(\varSigma )\), and let \(M\subset D\times \mathbb {R}\) be the submanifold of the product defined using the curve \(\gamma \), as in Sect. 3. Let g be the metric on D, while \(\tilde{g}\) is the metric induced by the product metric \(\bar{g}=g+dt^2\) on M, and h is the metric induced (from g or \(\tilde{g}\)) on a slice \(M\cap (D\times \{t\})\), which is diffeomorphic to \(\mathbb {S}^{n-1}\) or \(\varSigma \times \mathbb {S}^{q-1}\), respectively. We can view h as an r-dependent metric, or a t-dependent metric, via the relation \((r, t)\in \gamma \). We can write \(\tilde{g}= h + ds^2\), as the s-direction along the appropriate \(\gamma _\beta \) (see Sect. 3) is tangent to M; similarly, the metric on D can be written as \(g=h+dr^2\), where \((r, t)\in \gamma \). For \(\tilde{f}\) a smooth function on \(D\times \mathbb {R}\), we can restrict \(\tilde{f}\) to M (and compute with respect to \(\tilde{g}\)), or we can restrict to a slice \(D\times \{ t\}\) (and compute with respect to g); we do not introduce new notation for these two restrictions of \(\tilde{f}\).

We show details for the computation of \(\varDelta _{\tilde{g}} \tilde{f}\) and \(|\tilde{\nabla } \tilde{f}|_{\tilde{g}}^2\) on \((M,\tilde{g})\). We first compute the following Christoffel symbols:

$$\begin{aligned} \tilde{\varGamma }_{ss}^\ell= & {} \frac{1}{2} \tilde{g}^{\ell m} \left( \tilde{g}_{sm,s} + \tilde{g}_{sm,s} - \tilde{g}_{ss,m}\right) = 0\\ \tilde{\varGamma }_{ij}^s= & {} \frac{1}{2} \tilde{g}^{sm} \left( \tilde{g}_{im,j} + \tilde{g}_{jm,i} - \tilde{g}_{ij,m}\right) = -\frac{1}{2} \tilde{g}^{ss} \tilde{g}_{ij,s} = -\frac{1}{2} h_{ij,s}, \end{aligned}$$

and similarly \(\tilde{\varGamma }^s_{sj}=0\).

We note \(\varDelta _{\tilde{g}} \tilde{f} = \tilde{g}^{ij}\left( \tilde{f}_{,ij} - \tilde{\varGamma }_{ij}^k \tilde{f}_{,k}\right) = \varDelta _{h} \tilde{f} - h^{ij} \tilde{\varGamma }_{ij}^s \tilde{f}_{,s} + \tilde{f}_{,ss} - \tilde{\varGamma }_{ss}^k \tilde{f}_k\), so that

$$\begin{aligned} \varDelta _{\tilde{g}} \tilde{f} = \varDelta _{h} \tilde{f} + \frac{1}{2} h^{ij} h_{ij,s} \tilde{f}_{,s} + \tilde{f}_{,ss}. \end{aligned}$$
(27)

Thus we also have

$$\begin{aligned} \varDelta _{g} \tilde{f}= & {} \varDelta _{h} \tilde{f} + \tilde{f}_{,rr} + \frac{1}{2} h^{ij} h_{ij,r} \tilde{f}_{,r} . \end{aligned}$$
(28)

Recall that \(\phi \) is the angle between the unit normal \(\nu \) to M and \(\frac{\partial }{\partial t}\), so that \(\frac{\partial }{\partial s} = -\cos \phi \frac{\partial }{\partial r} + \sin \phi \frac{\partial }{\partial t}\). Using (27) and (28), we have

$$\begin{aligned} \varDelta _{\tilde{g}} \tilde{f}= & {} \varDelta _{g} \tilde{f} - f_{,rr} - \frac{1}{2} h^{ij} h_{ij,r} \tilde{f}_{,r} + \frac{1}{2} h^{ij} h_{ij,s} \tilde{f}_{,s} + \tilde{f}_{,ss} \nonumber \\= & {} \varDelta _{g} \tilde{f} - \tilde{f}_{,rr} - \frac{1}{2} h^{ij} h_{ij,r} \tilde{f}_{,r} - \frac{1}{2} h^{ij} h_{ij,r} \cos \phi \left( -\cos \phi \tilde{f}_{,r} + \sin \phi \tilde{f}_{,t}\right) \nonumber \\&+\, \left( -\cos \phi \tilde{f}_{,r} + \sin \phi \tilde{f}_{,t}\right) _{,s} \nonumber \\= & {} \varDelta _{g} \tilde{f}- \tilde{f}_{,rr} - \frac{1}{2} h^{ij} h_{ij,r} \tilde{f}_{,r} - \frac{1}{2} h^{ij} h_{ij,r} \cos \phi \left( -\cos \phi \tilde{f}_{,r} + \sin \phi \tilde{f}_{,t}\right) \nonumber \\&+\, \sin \phi \frac{d\phi }{ds} \tilde{f}_{,r} - \cos \phi \left( -\cos \phi \tilde{f}_{,rr} + \sin \phi \tilde{f}_{,rt} \right) + \cos \phi \frac{d\phi }{ds} \tilde{f}_{,t} \nonumber \\&+\, \sin \phi \left( -\cos \phi \tilde{f}_{,tr} + \sin \phi \tilde{f}_{,tt}\right) \nonumber \\= & {} \varDelta _{ g} \tilde{f} + \sin ^2(\phi ) \left( \tilde{f}_{,tt} - \tilde{f}_{,rr}\right) + \tilde{f}_{,t}\left( -\frac{1}{2} h^{ij} h_{ij,r} \cos \phi \sin \phi +\, \cos \phi \frac{d\phi }{ds}\right) \nonumber \\&+\, \tilde{f}_{,r}\left( -\frac{1}{2} \sin ^2( \phi ) h^{ij} h_{ij,r}+ \sin \phi \frac{d\phi }{ds}\right) - 2\sin \phi \cos \phi \tilde{f}_{,tr}. \end{aligned}$$
(29)

We now compute \(|\tilde{\nabla } \tilde{f}|_{\tilde{g}}^2\). Note \(\tilde{\nabla }^\ell \tilde{f} = \tilde{g}^{\ell k} \tilde{f}_{,k}= h^{\ell k} \tilde{f}_{,k}+ \delta ^{\ell }_s \tilde{f}_{,s}\), while \({\nabla }^\ell \tilde{f} = {g}^{\ell k} \tilde{f}_k = h^{\ell k} \tilde{f}_k + \delta ^{\ell }_r \tilde{f}_{,r}\). Let Y be the vector field tangent to the level sets of r given in components by \(Y^\ell = h^{\ell k} \tilde{f}_{,k}\) . Then \(\tilde{\nabla } \tilde{f} = Y + \tilde{f}_{,s} \frac{\partial }{\partial s}\) and \(\nabla \tilde{f} = Y + \tilde{f}_{,r}\frac{\partial }{\partial r}\) are \(\tilde{g}\)-orthogonal and g-orthogonal decompositions, respectively. Therefore,

$$\begin{aligned} |\tilde{\nabla } \tilde{f}|_{\tilde{g}}^2= & {} |Y|_{\tilde{g}}^2+ (\tilde{f}_{,s})^2= |Y|_h^2 + (\tilde{f}_{,s})^2= |{\nabla } \tilde{f}|_{ g}^2+(\tilde{f}_{,s})^2-(\tilde{f}_{,r})^2 \nonumber \\= & {} |{\nabla } \tilde{f}|_{ g}^2 + 2 \sin \phi \cos \phi \tilde{f}_{,r} \tilde{f}_{,t} + \sin ^2( \phi ) \left[ (\tilde{f}_{,t})^2 - (\tilde{f}_{,r})^2\right] . \end{aligned}$$
(30)

1.2 Metric Estimates for Tubes

Lemma 4

Suppose \(\varSigma \subset (X,g)\) is a closed submanifold of codimension \(q\ge 3\), with normal bundle \(N(\varSigma )\). Take \(r_0>0\) so that for \(0<r<r_0\), there is a regular tubular neighborhood \(\mathcal {N}_r(\varSigma )= \{ \exp _p(v): p\in \varSigma ,\; v\in N_p\varSigma ,\; |v|_g< r\}\) about \(\varSigma \) in M. Under a local trivialization of the normal bundle over \(U\subset \varSigma \) using smooth orthonormal sections \(\nu _1, \ldots , \nu _q\) of N(U), we have a diffeomorphism \(\varPsi : \mathcal {N}_{r_0}(U) \rightarrow U \times \mathbb {D}^q(r_0)\). Using a covering of \(\varSigma \) by such open sets U, the pullback metrics \(\varPsi ^*(g|_{T\varSigma } \oplus g_{\mathrm {eucl}})\) patch together to give a metric \(\mathring{g}\) on \(\mathcal {N}_{r_0}(\varSigma )\). For \(\epsilon >0\) sufficiently small, there is a smooth homotopy through metrics with positive scalar curvature on \(\partial \mathcal {N}_{\epsilon }( \varSigma )\) from the metric \(h_{\epsilon }\) induced from g to the metric \(\mathring{h}_{\epsilon }\) induced from \(\mathring{g}\).

Proof

We begin by setting up notation, and record the expansion of the metric \(h_{\epsilon }\) on \(\partial \mathcal {N}_{\epsilon }(\varSigma )\). We first choose adapted coordinates \((F=\varPsi ^{-1},(x^1, \ldots x^q, y^1, \ldots , y^\ell ))\) on \(\mathcal {N}_{r_0}(U)\), where \((f,y=(y^1, \ldots , y^\ell ))\) are coordinates for an open set \(U\subset \varSigma \), say \(f(y)\in U\), and then \(F(x,y)= \exp _{f(y)}(x^i \nu _i)\), and \(\sum _{i=1}^q (x^i)^2 <r_0\). For instance, we could let f be given by normal coordinates at \(p\in U\), which we assume here. If we let \(z\mapsto \varTheta (z)\in \mathbb {S}^{q-1}\subset \mathbb {R}^q\) be a local parameterization of the (unit) sphere, then we can obtain local coordinates on \(\partial \mathcal {N}_r(\varSigma )\) by \(G_r(y,z)= F(r\varTheta (z), y)\).

We want to expand the metric \(h_{\epsilon }\) on \(\partial \mathcal {N}_{\epsilon }(\varSigma )\) for small \(\epsilon \). Fix \(p\in \varSigma \), and let f be normal coordinates on \(\varSigma \) centered at \(p=f(0)\). The coordinate frame induced by F evaluates at p to \(\{E_1, \ldots , E_{n}\}\), where \(E_k= F_*\Big |_{(0,0)}\left( \frac{\partial }{\partial x^k}\right) ,\;k\in \{ 1, \ldots , q\}\), and \(E_{q+j}= F_*\Big |_{(0,0)}\left( \frac{\partial }{\partial y^j}\right) = f_*\Big |_{0}\left( \frac{\partial }{\partial y^j}\right) ,\;j\in \{ 1, \ldots \ell \}\). Note that \(E_k= \nu _k\) at p, for \(k=1, \ldots , q\). By use of normal coordinates at p, we see that \((E_1, \ldots , E_{n})\) is an orthonormal frame for \(T_p X\). We let \(Z_j\Big |_{G_{\epsilon }(y,z)}= (G_{\epsilon })_*\Big |_{(y,z)}\left( \frac{\partial }{\partial z^j}\right) \) for \(j\in \{1, \ldots , q-1\}\), and \(Y_{a}\Big |_{G_{\epsilon }(y,z)}=( G_{\epsilon })_*\Big |_{(y,z)}\left( \frac{\partial }{\partial y^a}\right) \) for \(a\in \{1, \ldots , \ell \}\). In what follows below, we let \(i, j \in \{ 1, \ldots , q-1\}\), while \(a, b, c, d \in \{ 1, \ldots , \ell \}\).

Then by [19, cf. Proposition 2.1], we have the following at \(q=G_{\epsilon }( 0,z)\) where the partial derivatives are evaluated at z:

$$\begin{aligned} h_{\epsilon }(Y_a, Y_b)= & {} g(Y_a, Y_b) =\delta _{ab} + O(\epsilon )\\ \epsilon ^{-1}h_{\epsilon }(Y_a, Z_j)= & {} \epsilon ^{-1} g(Y_a, Z_j)= O(\epsilon ^2)\\ \epsilon ^{-2} h_{\epsilon }(Z_i, Z_j)= & {} \epsilon ^{-2} g(Z_i, Z_j)= \sum _{k=1}^q \frac{\partial \varTheta ^k}{\partial z^i} \frac{\partial \varTheta ^k}{\partial z^j} +O(\epsilon ^2). \end{aligned}$$

These give the components of the metric \(h_{\epsilon }\) in the given chart.

We compare this to the metric \(\mathring{h}_{\epsilon }\) on \(\partial \mathcal {N}_{\epsilon }(\varSigma )\) induced locally by the product metric on \(\varSigma \times \mathbb {S}^{q-1}\), where the spherical factor has the round metric \(\epsilon ^2 \mathring{\sigma } \) of radius \(\epsilon \). The metric \(\mathring{h}_{\epsilon }\) is given at \(q=G_{\epsilon }(0,z)\) by

$$\begin{aligned} \mathring{h}_{\epsilon }(Y_a, Y_b)= & {} \delta _{ab} \\ \epsilon ^{-1}\mathring{h}_{\epsilon }(Y_a, Z_j)= & {} 0\\ \epsilon ^{-2} \mathring{h}_{\epsilon }(Z_i, Z_j)= & {} \mathring{g}\left( \frac{\partial \varTheta ^k}{\partial z^i}E_k, \frac{\partial \varTheta ^{\ell }}{\partial z^j}E_{\ell }\right) = \sum _{k=1}^q \frac{\partial \varTheta ^k}{\partial z^i} \frac{\partial \varTheta ^k}{\partial z^j} . \end{aligned}$$

This implies that the difference is given by

$$\begin{aligned} h_{\epsilon }-\mathring{h}_{\epsilon }=A_{ab}(y,z) dy^a dy^b + 2B_{cj}(y,z) dy^c dz^j + C_{ij} (y,z) dz^i dz^j, \end{aligned}$$

where \(A_{ab}=A_{ba},\;C_{ij}=C_{ji}\), and where for one-forms \(\theta \) and \(\eta \), we let \(\theta \eta = \frac{1}{2} (\theta \otimes \eta + \eta \otimes \theta )\). From the above, we have \(A_{ab}=O(\epsilon ),\;B_{cj}= O(\epsilon ^3)\) and \(C_{ij}=O(\epsilon ^4)\); partial derivatives in \(y^a\) or \(z^i\) of these component functions have the same error estimates. (We can cover \(\varSigma \) by finitely many such charts, and obtain uniform error estimates.) This implies that the raised metrics (each respectively raised with respect to itself) are related by

$$\begin{aligned} h_{\epsilon }^{\sharp }(y,z)= \mathring{h}_{\epsilon }^{\sharp }(y,z)+ \widehat{A}^{ab}(y,z) \partial _{y^a} \partial _{y^b} + 2\widehat{B}^{cj}(y,z) \partial _{y^c} \partial _{z^j} + \widehat{C}^{ij}(y,z) \partial _{z^i} \partial _{z^j}, \end{aligned}$$

where \(\widehat{A}^{ab}=\widehat{A}^{ba},\;\widehat{C}^{ij}=\widehat{C}^{ji}\) and for vectors X and \(Y,\;X Y= \frac{1}{2} (X\otimes Y+ Y\otimes X)\). Here we have \(\widehat{A}^{ab}=O(\epsilon ),\;\widehat{B}^{cj}=O(\epsilon )\) and \(\widehat{C}^{ij}=O(1)\), with the same estimates holding for the \(y^a\) and \(z^i\) partial derivatives, i.e., for any multi-indices \(\alpha \) and \(\beta \), there is a constant \(C=C_{\alpha \beta }\) so that for all \(a,\;b,\;|\partial ^{\alpha }_y \partial ^{\beta }_z \widehat{A}^{ab}|\le C \epsilon \), and analogously for the other components.

We wish to compute the scalar curvature, for which we record the following estimates for the Christoffel symbols \(\varGamma ^{\mu }_{\kappa \lambda }\) of \(h_{\epsilon }\). For the summation convention, we let \(a, b, c, d \in \{ 1, \ldots , \ell \}\) correspond to directions along \(\varSigma \), whereas \(s, t, u,v \in \{ \ell +1, \ldots , \ell +q-1=n-1\}\) correspond to spherical directions (so \(s=j+\ell \) corresponds to \(z^j\)), and Greek letters denote elements of \(\{ 1, \ldots , n-1\}\). We let \(\mathring{\varGamma }^{\mu }_{\kappa \lambda }\) be the Christoffel symbols for \(\mathring{g}\), so that \(\mathring{\varGamma }^{a}_{bc}\) are the Christoffel symbols for the metric h on \(\varSigma \) in the coordinates y, whereas \(\mathring{\varGamma }^{s}_{tu}\) are the Christoffel symbols of the round metric \(\mathring{\sigma }\) in the coordinates z, where we note that the Christoffel symbols are invariant under constant metric rescaling. In the equations below, \(\Lambda =O(\epsilon ^m)\) for a function \(\Lambda \) on (an open subset of) \(\varSigma \times \mathbb {S}^{q-1}\times (0, r_0)\) pulled back from \(\mathcal {N}_{r_0} (\varSigma ) \setminus \varSigma \) means there exist \(C>0\) so that \(|\Lambda |\le C \epsilon ^m\), and for any index a or \(j,\;|\partial _{y^a} \Lambda |\le C\epsilon ^m,\;|\partial _{z^j} \Lambda | \le C \epsilon ^m\) (and likewise for higher derivatives).

Lemma 5

For \(a, b, c \in \{ 1, \ldots , \ell \}\) and \(s, t, u \in \{ \ell +1, \ldots , n-1\}\),

$$\begin{aligned} \varGamma ^{s}_{tu}= \mathring{\varGamma }^{s}_{tu} + O(\epsilon ^2),&\qquad&\varGamma ^a_{bc}=\mathring{\varGamma }^a_{bc}+O(\epsilon )\\ \varGamma ^{a}_{tu}= O(\epsilon ^3),&\varGamma ^a_{tb} = O(\epsilon )\\ \varGamma ^s_{ta}= O(\epsilon ),&\varGamma ^{s}_{bc}= O(\epsilon ^{-1}). \end{aligned}$$

Proof

The proof is a simple calculation, using the above metric estimates. We pay particular attention to terms involving \((\mathring{h}_{\epsilon })^{st}=\epsilon ^{-2} \mathring{\sigma }^{st}\). We have

$$\begin{aligned} \varGamma ^{s}_{tu}= & {} \frac{1}{2} (h_{\epsilon })^{sv} \left( (h_{\epsilon })_{vu,t} + (h_{\epsilon })_{tv,u}-(h_{\epsilon })_{tu,v}\right) + \frac{1}{2} (h_{\epsilon })^{sa}((h_{\epsilon })_{au,t}+(h_{\epsilon })_{ta,u}-(h_{\epsilon })_{tu,a})\\= & {} \mathring{\varGamma }^s_{tu}+ \frac{1}{2} (\mathring{h}_{\epsilon })^{sv} \left( C_{vu,t} + C_{tv,u}-C_{tu,v}\right) +\frac{1}{2} \widehat{C}^{sv} \left( (h_{\epsilon })_{vu,t} + (h_{\epsilon })_{tv,u}-(h_{\epsilon })_{tu,v}\right) \\&+\, \frac{1}{2} \widehat{B}^{sa} \left( (h_{\epsilon })_{au,t}+(h_{\epsilon })_{ta,u}-(h_{\epsilon })_{tu,a}\right) \\= & {} \mathring{\varGamma }^s_{tu} + O(\epsilon ^2). \end{aligned}$$

Similarly, we see

$$\begin{aligned} \varGamma ^{a}_{bc}= & {} \frac{1}{2} (h_{\epsilon })^{as} \left( (h_{\epsilon })_{sc,b} + (h_{\epsilon })_{bs,c}-(h_{\epsilon })_{bc,s}\right) + \frac{1}{2} (h_{\epsilon })^{ad}((h_{\epsilon })_{dc,b}+(h_{\epsilon })_{bd,c}-(h_{\epsilon })_{bc,d})\\= & {} \frac{1}{2} \widehat{B}^{as} \left( (h_{\epsilon })_{sc,b} + (h_{\epsilon })_{bs,c}-(h_{\epsilon })_{bc,s}\right) + \mathring{\varGamma }^a_{bc} + \frac{1}{2} (\mathring{h}_{\epsilon })^{ad}\left( A_{dc,b} + A_{bd,c}-A_{bc,d}\right) \\&+\, \frac{1}{2} \widehat{A}^{ad} \left( (h_{\epsilon })_{dc,b}+(h_{\epsilon })_{bd,c}-(h_{\epsilon })_{bc,d}\right) \\= & {} \mathring{\varGamma }^a_{bc} + O(\epsilon ). \end{aligned}$$

We compute similarly with the other terms:

$$\begin{aligned} \varGamma ^{a}_{tu}= & {} \frac{1}{2} (h_{\epsilon })^{ab} \left( (h_{\epsilon })_{bu,t} + (h_{\epsilon })_{tb,u}-(h_{\epsilon })_{tu,b}\right) + \frac{1}{2} (h_{\epsilon })^{as}\left( (h_{\epsilon })_{su,t}+(h_{\epsilon })_{ts,u}-(h_{\epsilon })_{tu,s}\right) \\= & {} \frac{1}{2} (h_{\epsilon })^{ab} \left( B_{bu,t}+B_{tb,u}-C_{tu,b}\right) + \frac{1}{2} \hat{B}^{as} \left( (h_{\epsilon })_{su,t}+(h_{\epsilon })_{ts,u}-(h_{\epsilon })_{tu,s}\right) \\= & {} O(\epsilon ^3) \\ \varGamma ^{a}_{tb}= & {} \frac{1}{2} (h_{\epsilon })^{ac} \left( (h_{\epsilon })_{cb,t}+(h_{\epsilon })_{tc,b}-(h_{\epsilon })_{tb,c}\right) + \frac{1}{2} (h_{\epsilon })^{as} \left( (h_{\epsilon })_{sb,t}+(h_{\epsilon })_{ts,b}-(h_{\epsilon })_{tb,s} \right) \\= & {} \frac{1}{2} (h_{\epsilon })^{ac} \left( A_{cb,t}+B_{tc,b}-B_{tb,c}\right) + \frac{1}{2} \widehat{B}^{as} \left( B_{sb,t}+C_{ts,b}-B_{tb,s} \right) \\= & {} O(\epsilon )\\ \varGamma ^{s}_{ta}= & {} \frac{1}{2} (h_{\epsilon })^{sb} \left( (h_{\epsilon })_{ba,t}+(h_{\epsilon })_{tb,a}-(h_{\epsilon })_{ta,b}\right) + \frac{1}{2} (h_{\epsilon })^{su}\left( (h_{\epsilon })_{ua,t}+(h_{\epsilon })_{tu,a}-(h_{\epsilon })_{ta,u}\right) \\= & {} \frac{1}{2} \widehat{B}^{sb} \left( A_{ba,t}+ B_{tb,a}-B_{ta.b}\right) + \frac{1}{2} \left( (\mathring{h}_{\epsilon })^{su}+ \widehat{C}^{su} \right) \left( B_{ua,t}+C_{tu,a}-B_{ta,u}\right) \\= & {} O(\epsilon )\\ \varGamma ^{s}_{bc}= & {} \frac{1}{2} (h_{\epsilon })^{sa} \left( (h_{\epsilon })_{ac,b}+(h_{\epsilon })_{ba,c}-(h_{\epsilon })_{bc,a}\right) + \frac{1}{2} (h_{\epsilon })^{su}\left( (h_{\epsilon })_{uc,b}+(h_{\epsilon })_{bu,c}-(h_{\epsilon })_{bc,u}\right) \\= & {} \frac{1}{2} \widehat{B}^{sa} \left( (h_{\epsilon })_{ac,b}+(h_{\epsilon })_{ba,c}-(h_{\epsilon })_{bc,a}\right) \\&+\,\frac{1}{2} (\mathring{h}^{su}+ \widehat{C}^{su} ) \left( (\mathring{h}_{\epsilon })_{uc,b}+B_{uc,b}+(\mathring{h}_{\epsilon })_{bu,c}+B_{bu,c}-(\mathring{h}_{\epsilon })_{bc,u}-A_{bc,u}\right) \\= & {} O(\epsilon )+ \frac{1}{2} \left( \epsilon ^{-2}\mathring{\sigma }^{su}+ \widehat{C}^{su} \right) \left( B_{uc,b}+B_{bu,c}-A_{bc,u}\right) \\= & {} O(\epsilon ^{-1}). \end{aligned}$$

\(\square \)

Using the above lemma, a straightforward expansion of the scalar curvature in terms of Christoffel symbols, paying special attention to terms involving \(\varGamma ^s_{bc}=O(\epsilon ^{-1})\) and \((\mathring{h}_{\epsilon })^{st}= \epsilon ^{-2} \mathring{\sigma }^{st}\), now shows that the scalar curvature satisfies the estimate

$$\begin{aligned} R(h_\epsilon )= R(\mathring{h}_{\epsilon })+ O(\epsilon ^{-1}) = R(\mathring{h}_{\epsilon })(1+ O(\epsilon ))\ge C \epsilon ^{-2}. \end{aligned}$$

For \(\epsilon \) sufficiently small, then, not only is \(R(h_{\epsilon })\) positive, but the above argument applies to the metric \(h_{\epsilon }+t (\mathring{h}_{\epsilon }- h_{\epsilon })= \mathring{h}_{\epsilon }+ (1-t)(h_{\epsilon }- \mathring{h}_{\epsilon })\), from we can conclude \(R(h_{\epsilon }+t (\mathring{h}_{\epsilon }- h_{\epsilon }))>0\) for all \(t\in [0,1]\). Thus we can isotope \(h_{\epsilon }\) to \(\mathring{h}_{\epsilon }\) through metrics of positive scalar curvature.

We have the following as an immediate corollary of the proof.

Corollary 5

Let \(\mathcal {N}_{r_0}(\varSigma )\subset (X,g)\) be a regular tubular neighborhood of a closed, embedded submanifold \(\varSigma \) of codimension \(q\ge 3\). For \(0<\epsilon <r_0\), let \(g_\epsilon \) be the metric induced on \(\partial \mathcal {N}_{\epsilon }(\varSigma )\) from g, and let \(\mathring{h}_{\epsilon }\) be the metric on \(\partial \mathcal {N}_{\epsilon } (\varSigma )\) obtained by pullback of \(\gamma \oplus \epsilon ^2 \mathring{\sigma }\) via the exponential map and a covering of local trivializations of the normal bundle as above, where \(\mathring{\sigma }\) is the standard unit round metric on \(\mathbb {S}^{q-1}\), and \(\gamma \) is any metric on \(\varSigma \). By choosing \(\epsilon >0\) sufficiently small, one can isotope \(g_{\epsilon }\) to \(\mathring{h}_{\epsilon }\) through metrics of positive scalar curvature.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Abedin, F., Corvino, J. On the P-Scalar Curvature. J Geom Anal 27, 1589–1623 (2017). https://doi.org/10.1007/s12220-016-9732-9

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12220-016-9732-9

Keywords

Mathematics Subject Classification

Navigation