Skip to main content
Log in

Effective Results on Uniformization and Intrinsic GCM Spheres in Perturbations of Kerr

  • Manuscript
  • Published:
Annals of PDE Aims and scope Submit manuscript

Abstract

This is a follow-up of our paper (Klainerman and Szeftel in Construction of GCM spheres in perturbations of Kerr, Accepted for publication in Annals of PDE) on the construction of general covariant modulated (GCM) spheres in perturbations of Kerr, which we expect to play a central role in establishing their nonlinear stability. We reformulate the main results of that paper using a canonical definition of \(\ell =1\) modes on a 2-sphere embedded in a \(1+3\) vacuum manifold. This is based on a new, effective, version of the classical uniformization theorem which allows us to define such modes and prove their stability for spheres with comparable metrics. The reformulation allows us to prove a second, intrinsic, existence theorem for GCM spheres, expressed purely in terms of geometric quantities defined on it. A natural definition of angular momentum for such GCM spheres is also introduced, which we expect to play a key role in determining the final angular momentum for general perturbations of Kerr.

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.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Notes

  1. Such as formation of black holes, stability of multiple black holes, final state conjecture,...

  2. As explained in Remark 1.3 of [17], this is not strictly necessary. Any other foliation satisfying comparable asymptotic assumptions would also work.

  3. Recall that on the standard sphere  \(\mathbb {S}^2\), in spherical coordinates \(({\theta }, {\varphi })\), these are \(J^{(0, \mathbb {S}^2)}=\cos {\theta }\), \(J^{(+,\mathbb {S}^2)}=\sin {\theta }\cos {\varphi }\), \(J^{(-,\mathbb {S}^2)}=\sin {\theta }\sin {\varphi }\).

  4. See the introduction in [17] for further explanations.

  5. Compatible with small perturbations of Kerr.

  6. Note that the GCM conditions (1.9) require a choice of \(\ell =1\) modes on \(\mathbf{S}\), see Remark  1.2.

  7. We thank A. Chang for bringing this works to our attention. The precise formulation given in Theorem  3.1 does not seem to appear in the literature we are aware of.

  8. We thank C. De Lellis for bringing this paper to our attention.

  9. i.e. all the solutions are of the form \((\Phi \circ O, u\circ O)\) for \(O\in O(3)\).

  10. Up to a rotation of \(\mathbb {S}^2\).

  11. Up to a rotation of \(\mathbb {S}^2\).

  12. The only case where the ambiguity cannot be removed is when (1.24) holds for any one choice of canonical modes, i.e. \((\text{ curl }\,\beta ^\mathbf{S})_{\ell =1}=0\). In that case, we set \(a^\mathbf{S}=0\) which is consistent with (1.25).

  13. Note that in a Kerr space \({{\mathcal {K}}}(a, m)\), relative to a geodesic foliation normalized on \({{\mathcal {I}}}^+\), we have \( \int _\mathbf{S}\text{ curl }\,^\mathbf{S}\beta ^\mathbf{S}J^{(\pm , \mathbf{S})}=0\) and \( \int _\mathbf{S}\text{ curl }\,^\mathbf{S}\beta ^\mathbf{S}J^{(0, \mathbf{S})} =\frac{8\pi a m}{ (r^\mathbf{S})^3}+O\left( \frac{ma^2}{(r^\mathbf{S})^4}\right) \).

  14. This follows immediately from writing \(\Phi ^\# \gamma _0= e^{2u} \gamma _0\) in matrix form, by evaluating on an orthonormal frame, and then taking the absolute value of the determinant on both sides. Also, recall that \(|\det d\Phi |\) is an intrinsic scalar on \(\mathbb {S}^2\), i.e. it does not depend on the particular choice of orthonormal frame.

  15. To the best of our knowledge the condition \(\int _{\mathbb {S}^2} x e^{2u}=0\) appears first in [1].

  16. i.e. all the solutions are of the form \((\Phi \circ O, u\circ O)\) for \(O\in O(3)\). In particular, recall that if u is centered, then so is \(u\circ O\) for \(O\in O(3)\), see Remark  2.9.

  17. Recall from Remark 2.9 that if u belongs to \({{\mathcal {S}}}\), then \(u\circ O\) also belongs to \({{\mathcal {S}}}\) for any \(O\in O(3)\).

  18. Note that the kernel of \(\Delta _0 + 2\) is given by the \(\ell =1\) spherical harmonics, and that \(\int u x\) corresponds to the projection on these spherical harmonics.

  19. i.e. all the solutions are of the form \((\Phi \circ O, u\circ O)\) for \(O\in O(3)\).

  20. Note that, a priori, one would expect the right-hand side of (3.23) to be \(O(\epsilon )\). The fact that it is actually \(O(\epsilon ^2)\) is an application of Corollary 3.8 and (3.21), see the proof below.

  21. The result in [10] actually covers only \(p=2\). The case \(1<p<+\infty \) is in section 2.4 of S. Conti and B. Schweizer [9].

  22. The statement of Lemma 3.6 actually yields \(|\psi -I|\lesssim \delta \). Here we use the fact that \({{\mathcal {S}}}\) is invariant by O(3), see Remark 2.9, so that we may indeed assume, by composing with \(O^{-1}\in O(3)\), that \(|\psi -O^{-1}|\lesssim \delta \).

  23. More precisely, \(B_1\) and \(B_2\) of Theorem 4.6 are chosen to be euclidean balls centered on points of \(\mathbb {S}^2\) of radius given respectively by 1/4 and 1/2 so that both \(B_1\) and \(B_2\) are included in D and the union of the \(B_1\)’s includes \(D_1\).

  24. Given by Corollary 3.8.

  25. In particular, one can choose \(N=(0,0,1)\) and \(v=(1,0,0)\).

  26. Note that the kernel of \(\Delta _0 + 2\) is given by the \(\ell =1\) spherical harmonics, and that \(\int u x\) corresponds to the projection on these spherical harmonics.

  27. That is \(\mathbf{g}^{\alpha \beta }\partial _\alpha u\partial _\beta u =0\).

  28. In view of (5.11), we will often replace \(\Gamma _g\) by \(r^{-1} \Gamma _b\).

  29. That is the quantities on the left verify the same estimates as those for \(\Gamma _b\), respectively \(\Gamma _g\).

  30. The properties (5.14) of the scalar functions \(J^{(p)}\) are motivated by the fact that the \(\ell =1\) spherical harmonics on the standard sphere \(\mathbb {S}^2\), given by \(J^{(0, \mathbb {S}^2)}=\cos {\theta }, \, J^{(+, \mathbb {S}^2)}=\sin {\theta }\cos {\varphi }, \, J^{(-, \mathbb {S}^2)}=\sin {\theta }\sin {\varphi }\), satisfy (5.14) with \(\overset{\circ }{\epsilon }=0\). Note also that on \(\mathbb {S}^2\),

    $$\begin{aligned} \int _{{\mathbb {S}}^2}(\cos {\theta })^2=\int _{{\mathbb {S}}^2}(\sin {\theta }\cos {\varphi })^2=\int _{{\mathbb {S}}^2}(\sin {\theta }\sin {\varphi })^2=\frac{4\pi }{3}, \qquad |\mathbb {S}^2|=4\pi . \end{aligned}$$
  31. Note that while the Ricci coefficients \(\kappa ^\mathbf{S}, {\underline{\kappa }}^\mathbf{S}, {\widehat{\chi }}^\mathbf{S}, {\widehat{{\underline{\chi }}}}^\mathbf{S}, \zeta ^\mathbf{S}\) as well as all curvature components and mass aspect function \(\mu ^\mathbf{S}\) are well defined on \(\mathbf{S}\), this in not the case of \(\eta ^\mathbf{S}, {\underline{\eta }}^\mathbf{S}, \xi ^\mathbf{S}, {\underline{\xi }}^\mathbf{S}, \omega ^\mathbf{S}, {\underline{\omega }}^\mathbf{S}\) which require the derivatives of the frame in the \(e_3^\mathbf{S}\) and \(e_4^\mathbf{S}\) directions.

  32. Note that (6.2) implies (6.15) in view of (5.27).

  33. \(J^{(p,\overset{\circ }{ S})}\) denotes the canonical basis of \(\ell =1\) modes on \(\overset{\circ }{ S}\) corresponding to the effective uniformization map \((\overset{\circ }{\Phi }\,, \overset{\circ }{\phi }\,)\) for \(\overset{\circ }{ S}\) appearing in Definition  6.3.

  34. Recall that \(\delta _1\) is the smallness constant introduced in A1-Strong, see (7.1).

  35. Note that in a Kerr space \({{\mathcal {K}}}(a, m)\), relative to a geodesic foliation normalized on \({{\mathcal {I}}}^+\), we have \( \int _\mathbf{S}\text{ curl }\,^\mathbf{S}\beta ^\mathbf{S}J^{\pm , S}=0\) and \( \int _\mathbf{S}\text{ curl }\,^\mathbf{S}\beta ^\mathbf{S}J^{0, S} =\frac{8\pi a m}{ (r^\mathbf{S})^3}+O\left( \frac{ma^2}{ (r^\mathbf{S})^4}\right) \).

  36. Note that in Kerr, the axis corresponds to the points where \(J^{(0,\mathbf{S})}=\pm 1\).

  37. This relies in particular on the following analog of the estimate for \(\text{ div }\,\beta \) in the proof of Lemma 7.9

    $$\begin{aligned} \text{ curl }\,\beta = \frac{6a_0m_0\cos {\theta }}{r^5}+O(a_0^2m_0r^{-6}). \end{aligned}$$
  38. \({\widetilde{J}}^{(p)}\) is not assumed to be a canonical basis of \(\ell =1\) modes on \(\mathbf{S}\).

  39. The only step which requires a slight modification of the argument is when passing from (7.13) to (7.14) as \((\kappa -\kappa ^\mathbf{S})\Delta ^\mathbf{S}\text{ div }\,^\mathbf{S}{\underline{f}}\) induces the term \((\frac{2}{r}-\frac{2}{r^\mathbf{S}})\Delta ^\mathbf{S}\text{ div }\,^\mathbf{S}{\underline{f}}\) which would loose one more power of r than allowed. Instead, one should first divide (7.13) by \(\kappa \) and the proof then proceeds along the same lines.

References

  1. Aubin, T.: Meilleures constantes dans le théorème d’inclusion de Sobolev et un théorème de Fredholm non linéaire pour la transformation conforme de la courbure scalaire. J. Funct. Anal. 32, 148–174 (1979)

    Article  MATH  Google Scholar 

  2. Berger, M.: Geometry II. Universitext. Springer (1987)

  3. Chang, S.-Y.A., Yang, P.: Prescribing Gaussian curvature on \(S^2\). Acta Math. 159, 215–259 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  4. Chang, S.-Y.A., Yang, P.: A perturbation result on prescribing scalar curvature on \(S^n\). Duke Math. J. 64, 27–69 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  5. Chang, S.-Y.A.: The Moser-Trudinger inequality and applications to some problems in conformal geometry. IAS/Park City Mathematical series 65–125 (1996)

  6. Chen, P.-N., Wang, M.-T., Yau, S.-T.: Quasilocal angular momentum and center of mass in general relativity. Adv. Theor. Math. Phys. 20, 671–682 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  7. Christodoulou, D., Klainerman, S.: The Global Nonlinear Stability of the Minkowski Space. Princeton University Press, Princeton (1993)

    MATH  Google Scholar 

  8. Ciarlet, P.G.: On Korn’s inequality. Chin. Ann. Math. Ser B 31, 607–618 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  9. Conti, S., Schweizer, B.: Rigidity and Gamma convergence for solid–solid phase transition with \(SO(2)\) invariance. Commun. Pure Appl. Math. 59, 830–868 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  10. Friesecke, G., James, R., Müller, S.: A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Comm. Pure Appl. Math. 55, 1461–1506 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  11. Giorgi, E., Klainerman, S., Szeftel, J.: A general formalism for the stability of Kerr, arXiv:2002.02740

  12. Huisken, G., Yau, S.-T.: Definition of center of mass for isolated physical systems and unique foliations by stable spheres with constant mean curvature. Invent. Math. 124, 281–311 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. John, F.: Rotation and strain. Commun. Pure Appl. Math. 44, 391–413 (1961)

    Article  MathSciNet  MATH  Google Scholar 

  14. Jones, G., Singerman, D.: Complex Functions - An Algebraic and Geometric Viewpoint. Cambridge University Press, Cambridge (1987)

    Book  MATH  Google Scholar 

  15. Kirchheim, B., Spadaro, E., Székelyhidi, L.: Equidimensional isometric maps. Commun. Mah. Helvetici 90, 761–798 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  16. Klainerman, S., Szeftel, J.: Global Nonlinear Stability of Schwarzschild Spacetime Under Polarized Perturbations. Annals of Math Studies, vol. 210. Princeton University Press, Princeton (2020)

    Book  MATH  Google Scholar 

  17. Klainerman, S., Szeftel, J.: Construction of GCM spheres in perturbations of Kerr, Accepted for publication in Annals of PDE

  18. De Lellis, C.: Personal communication

  19. Onofri, E.: On the positivity of the effective action in a theory of random surfaces. Commun. Math. Phys. 86, 321–326 (1982)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Rizzi, A.: Angular momentum in general relativity: a new definition. Phys. Rev. Lett. 81(6), 1150–1153 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Szabados, L.B.: Quasi-local energy-momentum and angular momentum in general relativity. Living Rev. Relat. 12, 4 (2009)

    Article  ADS  MATH  Google Scholar 

Download references

Acknowledgements

The authors are grateful to A. Chang and C. De Lellis for very helpful suggestions and references in connection to our results on effective uniformization in sections 3 and 4.

The first author is supported by the NSF grant DMS 180841 as well as by the Simons grant 10011738. He would like to thank the Laboratoire Jacques-Louis Lions of Sorbonne Université and IHES for their hospitality during his many visits. The second author is supported by ERC grant ERC-2016 CoG 725589 EPGR.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sergiu Klainerman.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

With an appendix by Camillo De Lellis.

Appendix by Camillo De Lellis

Appendix by Camillo De Lellis

We denote by e the Euclidean metric on \({\mathbb {R}}^n\).

Theorem A.1

For every \(n\ge 2\) and every \(\alpha \in (0,1]\) there is a constant \(C = C(n,\alpha )\) with the following property. Let \(u\in H^1 (B_2, {\mathbb {R}}^n)\) be such that

$$\begin{aligned} \Vert u^\sharp e - e\Vert _{L^\infty (B_1)} \le \frac{1}{2} \end{aligned}$$
(A.1)

and

$$\begin{aligned} \det Du \ge 0 \qquad \text{ a.e. } \end{aligned}$$
(A.2)

Then there is \(A\in SO (n)\) such that

$$\begin{aligned} \Vert Du - A\Vert _{C^\alpha (B_1)} \le C \Vert u^\sharp e - e\Vert _{C^\alpha (B_2)}\, . \end{aligned}$$
(A.3)

Remark A.2

From now on, |A| will denote the Hilbert-Schmidt norm of the matrix \(A\in {\mathbb {R}}^{n\times n}\), namely \(|A|= \sqrt{\mathrm{tr}\, A^\top A}\) induced by the Hilbert-Schmidt scalar product \(\langle A, B\rangle := \mathrm{tr}\, A^\top B\). In particular, using standard coordinates in \(B_2\) and identifying \(u^\sharp e\) with the corresponding \(n\times n\) matrix, we have \(|Du|^2 = \mathrm{tr}\, u^\sharp e\). Under the assumption (A.1), we thus conclude immediately that \(\Vert Du\Vert _{L^\infty (B_2)} \le C (n)\), namely that the map u is Lipschitz.

Remark A.3

Note that an assumption like (A.2) is needed because it is easy to give examples of Lipschitz maps whose derivative belong to O(n) almost everywhere but which are not affine. In fact such maps are “abundant” in an appropriate sense: in particular they form a residual set in the space \(X:=\{u\in \mathrm{Lip}\, (B_2, {\mathbb {R}}^n): u^\sharp e \le e\}\) endowed with the the \(L^\infty \) distance, which makes X a compact metric space (cf. [15] for the latter and more subtle results).

The two authors of the paper asked me while preparing their manuscript whether I could provide a reference or a proof for Theorem A.1. While I felt that this should be a well-known “classical fact”, I was unable to find a reference for it. I therefore suggested a simple argument which reduces (A.3) to an important work of [10] which essentially handles a corresponding “\(L^2\)-estimate”. The reduction is given in this appendix. It uses some elementary facts from Linear Algebra (which are well known and I just include for the reader’s convenience) and a Morrey-type decay. In what follows we denote by \(\mathrm{Id}\) the identity matrix in \({\mathbb {R}}^{n\times n}\).

Lemma A.4

We have

$$\begin{aligned} \mathrm{dist}\,(A, SO (n)) = \mathrm{dist}\,(A, O(n)) \qquad \text{ for } \text{ all } A\in {\mathbb {R}}^{n\times n} \text{ with } \det A \ge 0 \end{aligned}$$
(A.4)

and

$$\begin{aligned} \mathrm{dist}\,(A, O(n)) \le |A^\top A - I\,d| \qquad \forall A\in {\mathbb {R}}^{n\times n}\, . \end{aligned}$$
(A.5)

Proof

In order to show (A.4) fix first an arbitrary matrix A with \(\det A \ge 0\). Recalling the polar decomposition of matrices there is a symmetric S and a \(O_1\in SO (n)\) such that \(A = O_1 S\). Next, recalling that every symmetric matrix is diagonalizable, there is \(O_2\in SO(n)\) such that \(A = O_1 O_2^\top D O_2\) for some diagonal matrix D. Next recall that if O is a diagonal matrix with an even number of entries equal to \(-1\) and the remaining equal to 1, then \(O\in SO(n)\). If one of the diagonal enties of D is zero, we can then assume without loss of generality that all enties of A are nonnegative. Otherwise, if no diagonal entry is 0, we can assume without loss of generality that they are all positive but at most 1. Since \(\det A>0\), we can exclude that one diagonal entry of D is negative and the others are all positive. Summarizing the arguments in the two cases, we can assume that all diagonal entries of D are nonnegative. Since \(\mathrm{dist}\,(A, O(n)) = \mathrm{dist}\,(OA, O(n))=\mathrm{dist}\,(AO, O(n))\) and \(\mathrm{dist}\,(A, SO(n))= \mathrm{dist}\,(OA, SO(n)) = \mathrm{dist}\,(AO, SO(n))\) for every \(O\in SO(n)\), we conclude that it suffices to prove (A.4) for a diagonal matrix A which has all nonnegative entries. Denote them by \(\lambda _1, \ldots , \lambda _n\). For any \(O\in O(n)\) we can then compute explicitely

$$\begin{aligned} |A - O|^2 = \sum _i \lambda _i^2 + n - 2\sum _i \lambda _i O_{ii}\, . \end{aligned}$$

Observe that \(-1\le O_{ii}\le 1\) because O is orthogonal. Since \(\lambda _i \ge 0\) for every i we then conclude \(|A-O|^2 \ge \sum _i \lambda _i^2 + n - 2 \sum _i \lambda _i = |A-I\,d|^2\). This however shows that \(\mathrm{dist}\,(A, O(n))^2 = |A-I\,d|^2 = \mathrm{dist}\,(A, SO(n))^2\).

As for (A.5) fix \(A\in {\mathbb {R}}^{n\times n}\) and let \(O\in O(n)\) be such that \(\mathrm{dist}\,(A, O(n)) = |A-O|\). Since both sides of the inequality take the same value for A and \(O^{-1} A\), we can assume that \(O = I\,d\). By the minimality condition of \(I\,d\) we must have that \(A-I\,d\) is orthogonal (in the Hilbert-Schmidt scalar product) to the tangent to O(n) at \(I\,d\), which is the space of skew-symmetric matrices. We therefore conclude that A is symmetric and, again applying the O(n) invariance of the inequality, we can assume w.l.o.g. that it is diagonal. Let \(\lambda _i= A_{ii}\) be the diagonal entries and observe that none of them can be negative: if \(\lambda _k <0\) then the matrix B which has \(B_{ij} =0\) for \(i\ne j\), \(B_{ii}=1\) for \(i\ne k\) and \(B_{kk}=-1\) satisfies \(B\in O(n)\) and \(|A-B|< |A-I\,d|\). (A.5) is thus reduced to proving

$$\begin{aligned} \sum _i (\lambda _i-1)^2 \le \sum _i (\lambda _i^2-1)^2\, \end{aligned}$$

under the assumption that \(\lambda _i \ge 0\) for every i. This is equivalent to prove \((x-1)^2 \le (x^2-1)^2= (x-1)^2 (x+1)^2\) for \(x\ge 0\), which is obvious. \(\square \)

Proof of Theorem A.1

Fix \(x\in B_1\) and let S(x) be the unique positive definite symmetric matrix such that \(S(x)^2 = u^\sharp e (x)\). Observe that, by (A.1), we have

$$\begin{aligned} |S(x)|+|S(x)^{-1}|\le C\, . \end{aligned}$$

Let v be the map \(v(y) := u (y) (S(x))^{-1}\) and observe further that

$$\begin{aligned} \mathrm{dist}\,(Dv (y), SO(n))&\le \mathrm{dist}\,(Dv (y), O(n)) \le |Dv^\top (y) Dv (y) - I\,d|\\&\le |S(x)^{-1} Du^\top (y) Du (y) S(x)^{-1} - I\,d|\\&= |S(x)^{-1} (Du^\top (y) Du (y) - S(y)^2)S(x)^{-1}|\\&= | S(x)^{-1} (u^\sharp e (y) - u^\sharp e (x)) S(x)^{-1}|\\&\le 4 [u^\sharp e]_{\alpha , B_2} |x-y|^\alpha = 4 [u^\sharp e - e]_{\alpha , B_2} |x-y|^\alpha \end{aligned}$$

(where we use the standard notation \([f]_{\alpha , \Omega }:= \sup \{\frac{|f(x)-f(y)|}{|x-y|^\alpha } : x,y\in \Omega \}\)). In particular, for every \(r\le 1\) we can apply the Friesecke-James-Müller inequality, namely [10, Theorem 3.1], to find a matrix \(A (x,r)\in SO(n)\) with the property that

(A.6)

(note that [10, Theorem 3.1] is stated for a general open set U in place of \(B_r (x)\), with a constant depending on U; however an obvious scaling argument shows that the constant is the same for balls of arbitrary radii). Recalling that

we conclude

Morrey’s estimate then implies that \(Du\in C^\alpha (B_1)\) and

$$\begin{aligned} {[}Du]_{\alpha , B_1} \le C [u^\sharp e - e]_{\alpha , B_2}\, . \end{aligned}$$
(A.7)

On the other hand, again by the Friesecke-James-Müller estimate, there is \(A\in SO (n)\) such that

(A.8)

(A.7) and (A.8) immediately imply (A.3). \(\square \)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Klainerman, S., Szeftel, J. Effective Results on Uniformization and Intrinsic GCM Spheres in Perturbations of Kerr. Ann. PDE 8, 18 (2022). https://doi.org/10.1007/s40818-022-00132-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40818-022-00132-7

Navigation