CONFORMAL CIRCLES AND LOCAL DIFFEOMORPHISMS
Abstract.
We study unparametrized conformal circles, or called conformal geodesics, study diffeomorphisms mapping conformal circles to conformal circles in pseudo-Riemannian conformal manifolds. We show that such local diffeomorphisms are conformal local diffeomorphisms. Our result extends the result of Yano and Tomonaga. We also present a holographic interpretation for our result on Poincaré-Einstein manifolds. The proofs take suitable variations of conformal circles.
1. Introduction
Riemannian geodesics, as fundamental geometric objects, are often considered when studying Riemannian structures. One classic problem concerning geodesics and Riemannian structures is the following: If there is a diffeomorphism that maps geodesics to geodesics, is it an isometry? The answer is negative due to affine transformations on Euclidean spaces. In general, one may need to further assume irreducible Riemannian manifolds for the diffeomorphism to be an isometry [13, 14]. The parallel problems for CR manifolds [5] and conformal manifolds [3, 24] are affirmative in some sense.
In the context of a pseudo-Riemannian conformal manifold with , a distinguished family of curves known as conformal circles or conformal geodesics emerges. These curves satisfy a third-order differential equation for non-null circles [1]. The derivation of these conformal circles is based on various perspectives of conformal manifolds, including Cartan geometry [10], the standard tractor bundle [2], and the Poincaré-Einstein manifold [9]. More detailed discussions about each viewpoint on conformal manifolds can be found in references such as [6, 8, 19, 25, 26]. Taking the unit sphere as a Riemannian conformal model, the conformal circles in the model are either straight lines or planar circles when viewed through stereographic projection onto Euclidean space [21]. In the context of the Poincaré ball , each circle on the boundary can be orthogonally extended to form a totally geodesic surface within . Based on the case of the conformally flat model, each conformal circle in a Riemannian conformal manifold can be formally extended to an asymptotically totally geodesic surface in the Poincaré-Einstein space by holographic construction [9] where the term holography comes from physics, e.g., [22]. We will use the term "holography" to mean the geometry from Poincaré-Einstein space .
In this article, we consider the parallel classic problem for conformal circles with a local diffeomorphism between pseudo-Riemannian conformal manifolds , with the same signature for and the problem in holographic settings . The problem for conformal circles in a Riemannian conformal manifold can be traced back to Carathéodory [3]. He considered a bijection on , which doesn’t need to be continuous, that maps straight lines (resp. circles) to straight lines (resp. circles), and he showed the bijection is a conformal transformation. Later on, K. Yano and Y. Tomonaga [23, 24] showed that an infinitesimal transformation is a conformal killing vector field if and only if it carries unit-speed conformal circles to unit-speed conformal circles. The equivalence problem is also discussed in Cartan geometry settings in terms of distinguished curves [7]. In §2, we review conformal circles and prove Theorem 2.3 for the equivalence problem of conformal circles. Our proof is motivated by the work from Yano and Tomonaga. In §3, we review Poincaré-Einstein manifolds and extend the results from Fine and Herfray [9] to indefinite signature. In §4, we prove Theorem 4.5 in holographic settings.
The author would like to thank his advisor, Jie Qing, for all discussions and illuminating advice.
2. Conformal Circles on Conformal Manifolds
In this section, we review definitions of conformal circles on a pseudo-Riemannian conformal manifold . As mentioned in the introduction, we prove Theorem 2.3 which extends the result of Yano and Tomonaga [23].
Let for and . The Levi-Civita connections and of respective and on sections of are related by [16]
where the Latin indices follow the convention in [18]. The is a conformally invariant tensor defined by
(1) |
where . We denote the Schouten tensor of and by and respectively for Their relations are
(2) |
When , we assume that (2) is a part of the structure on
Recall that Riemannian geodesics are the projections of integral curves of constant horizontal vector fields on linear frame bundle ([14] Chapter 3 Proposition 6.3); conformal circles can be defined in a similar way from Cartan geometry [10].
Definition 2.1.
Let and be the corresponding Levi-Civita connection. A conformal circle with respect to is a curve and a 1-form along such that they satisfy
(3) | ||||
(4) |
where the overdot of denotes the derivative with respect to . The equations are conformally invariant for conformal change if and .
Denote by for . We call a curve null if on and it is called non-null if it’s not null. Direct computation from (3) shows
(5) |
Therefore, if a conformal circle has a null velocity at some point, then it’s a null conformal circle. For completeness, we recall null pseudo-Riemannian geodesics are necessary and sufficient to be null conformal circles with some reparametrization [15, 21].
Assume a conformal circle is non-null. By solving in (3)
(6) |
one can have a third-order differential equation from (4). The third-order differential equation is equivalent to the system of the equations (3) and (4) if one defines the one form back.
Definition 2.2.
Given with corresponding Levi-Civita connection . A parametrized non-null conformal circle is defined to satisfy
(7) |
with initial conditions . The equation is invariant under conformal change .
Since the induced metric on a non-null curve is nondegenerate, we can consider the orthogonal decomposition of the pullback bundle of by [17]. We call satisfies the tangential (resp. normal) part of (7) if it is a solution of the equation that is the orthogonal projection of (7) to the tangent (resp. normal) bundle of . It is known [1] that any regular curve can be reparametrized to satisfy the tangential part of (7). The normal part of (7) is invariant under reparametrization of and it is only satisfied by non-null conformal circles. Since (7) is derived from (4), it’s convenient to introduce a vector field along an arbitrary curve
(8) |
where is a vector field along . If satisfies the right-hand side of (6) by descending index, then we denote the vector field by
For completeness, we recall another third-order differential equation for a non-null conformal circle . If is reparametrized so that it is of unit tangent velocity with respect to , then it satisfies [21]
The first line is the equation introduced by Yano [23].
Given a local diffeomorphism between pseudo-Riemannian conformal manifolds and . Assume both of the conformal classes have same signature . If is a conformal local diffeomorphism, it’s direct to see maps unparametrized conformal circles to unparametrized conformal circles. The converse direction is also true if the map preserves some nullity condition.
Theorem 2.3.
Let and be pseudo-Riemannian conformal manifolds with same signature . Assume a local diffeomorphism satisfying
-
(i)
is non-null (resp. null) non-null (resp. null) if ;
-
(ii)
sgn= sgn if .
If maps unparametrized non-null conformal circles to unparametrized non-null conformal circles, then is a conformal local diffeomorphism.
Proof.
Given . Let and . Choose a normal coordinate of centered at , . Since is a local diffeomorphism, we can identify the coordinate system near as . Let be a parametrized non-null conformal circle satisfying (7) with initial conditions , at where is the coordinate of with respect to . Since is an unparametrized non-null conformal circle, it satisfies the normal part of (7) with the given parameter , which is the following in the coordinate we chose.
(9) |
where is a vector field along defined from (8). In the following, we are considering for (9). Observe that (9) is a degree-two polynomial of with coefficients depending on the derivatives of and . It is because satisfies the third order differential equation (7) with respect to . Now let and where are fixed. The variable is an arbitrary number in an open interval containing . Because the non-null conformal circle equation is an autonomous ODE, (9) depends smoothly on . Based on the arguments we just made, we know (9) at is a polynomial of with degree two. Therefore, the coefficient of vanishes. After direct computation, the coefficient gives
(10) |
If , we get
(11) |
Assume the normal coordinate we chose is for and for . If and for , then from (11). If for , we then let and ; so, we have from (11). Therefore, the pullback metric is of the form
for some . If , then and are positive because is of the signature . If , then and can be both positive or both negative. Recalling that the preserves the nullity of null vectors, we know is null at with respect to when which implies . The sign of and is positive for since the sign of is the same as the sign of . ∎
Remark 2.4.
Remark 2.5.
The particular case for Theorem 2.3 has been studied in the literature if one assumes the to be a bijection (no need to be continuous) and with the standard Riemannian conformal structure [4, 12]. Note that the proof in [4, 12] needs the global property of conformal circles; that is, maps straight lines (resp. circles) to straight lines (resp. circles). However, by assuming additional regularity of in this paper, we only need the local condition of conformal circles, namely the conformal circle equation, to establish Theorem 2.3.
Remark 2.6.
Remark 2.7.
3. Holography Interpretation for Conformal Circles
In this section, we review Poincaré-Einstein manifolds and extend the results from Fine and Herfray [9] to indefinite signature.
Poincaré-Einstein manifold
Let be a pseudo-Riemannian conformal manifold with signature and . There exists a pseudo-Riemannian manifold with boundary [8]. The signature of is . Let be a defining function for , that is on the interior , on and on . Then, and satisfy
-
(i)
can be smoothly extended to be a metric on so that
-
(ii)
.
The pair is called a Poincaré-Einstein manifold for and the pair is called the conformal infinity to .
Let . There exists a unique defining function , called geodesic defining function, such that near . The function makes an identification between a neighborhood of in and a neighborhood of in . By the identification, is in normal form relative to ; that is, on where . The tensor is the Schouten tensor of for . When , is a symmetric two-tensor on satisfying and where is the scalar curvature of . By the pulling back of an even diffeomorphism between neighborhoods of which restricts to the identity map on , the normal forms relative to conformal related metrics are identical modulo .
Remark 3.1.
Note that the orders of above in the Ricci condition and in can be further refined to higher orders depending on dimension . See details in [8].
Remark 3.2.
For , the trace and the divergence conditions of are conformally invariant. ( [8], arguments after Theorem 3.7)
Surfaces in the Poincaré-Einstein Manifold
Let be a Poincaré-Einstein manifold in normal form relative to where is its corresponding geodesic defining function. Let be a non-null curve. The interval can be shrunk if necessary. Choose a local coordinate on an open set in containing for . The coordinate of is denoted by . If is an embedded surface orthogonal to with , then one can have an asymptotic isothermal coordinate of near ; that is, there is a diffeomorphism from to such that
(12) |
where and if and otherwise. In fact, to satisfy (12) for orthogonal to , the expansions of , and with respect to are in the following forms.
Proposition 3.3.
[9] Let be a Poincaré-Einstein metric and a surface as above. Then, the asymptotic isothermal coordinate in (12) satisfies
(13) |
and
(14) |
where the expansions of and are modulo , the denotes the square root of , the satisfies , the satisfies and which is the tangential part of (8). Note that is equal to the right-hand side of the -form in (6) by lowering index.
Since is pseudo-Riemannian for small , the tangent bundle has the orthogonal decomposition along near [17]. Let be a local orthonormal frame of the normal bundle of with respect to near . After direct computation, the Taylor expansions of with respect to are [9]
(15) | ||||
where is a family of sections of the normal bundle of in .
Proposition 3.4.
[9] Considering the projection of the second fundamental form of on with respect to , then, its asymptotic expansion is
(16) |
Due to Proposition 3.3, the asymptotic minimal condition of is equivalent to which is exactly the same as the -form in (6). The asymptotic totally geodesic condition is equivalently satisfied when , and being an unparametrized conformal circle.
Definition 3.5.
Let be an embedded surface orthogonal to so that is a non-null curve. It is called a proper surface if it is asymptotic totally geodesic where is the second fundamental form of with respect to .
4. Proof of Theorem
In this section, we consider a local diffeomorphism which smoothly extends a local diffeomorphism . We introduce the definition of asymptotic local isometry and cosider its local conditions. We also introduce an adapted coordinate for the proof of Theorem 4.5.
Let be a pseudo-Riemannian conformal manifold with same signature as and with a Poincaré-Einstein space . We keep as a coordinate system on an open set in .
Definition 4.1.
A local diffeomorphism is called an asymptotic local isometry if
(17) |
It’s useful to realize Definition 4.1 in terms of local coordinates. Let and be geodesic defining functions for and respectively. Identifying some neighborhoods of and to neighborhoods of and of respectively, a local diffeomorphism can be identified near and as a local diffeomorphism from to
(18) |
where , on and . Then, (17) is equivalent to
(19) |
where and . Since is a local diffeomorphism, we denote the coordinate of by with . In terms of the coordinates on and on , (19) is given by
(20) | ||||
(21) | ||||
(22) |
where we have used commas to express partial derivatives with respect to the coordinates on . At , (20) and (22) give
(23) |
If we choose and suitably such that is homothetic, that is constant, then (20)-(22) further imply at
(24) |
and at
(25) |
Conversely, if satisfies (23)-(25), then is an asymptotic local isometry.
Let be a surface orthogonal and its asymptotic isothermal coordinate (13). Since we are considering (23) to (25), it’s better to introduce a change of variables on , , where is equal to the right-hand side of in (13) modulo . Then, the expansions of and in terms of are
(26) |
where and remain the same conditions as in Proposition 3.3 and in Proposition 3.4. We call the adapted coordinate of .
Proposition 4.2.
Let be a local diffeomorphism in terms of (18). Assume maps non-null vectors in to non-null vectors in . Then, at if and only if is orthogonal to for any surface in that is orthogonal to and intersects along a non-null curve. In addition, if maps proper surfaces in to proper surfaces in , then at .
Proof.
Let be a non-null curve in a coordinate open set with the coordinate . Then, the adapted coordinate in (26) locally defines a surface orthogonal to with . Since is orthogonal to , then
(27) |
is orthogonal to where is the coordinate basis for the adapted coordinate . The orthogonal condition gives at
(28) |
Since is an arbitrary non-null vector at , we get at . Conversely, let be the adapted coordinate of . Projecting (27) to orthogonally, it gives the right-hand side of (28) which turns to be due to at .
If is a non-null conformal circle, then the formula of the adapted coordinate locally extends to a proper surface in . One can follow the same arguments just made to get at . ∎
As mentioned at the end of §2, the idea for proving Theorem 4.5 is to consider a suitable family of proper surfaces . Then, the dependence of in the second fundamental forms of proper surfaces may imply is an asymptotic local isometry where its local conditions are (23)-(25). However, recalling Proposition 3.3, Proposition 3.4 and the adapted coordinate (26), a proper surface is characterized by , and being an unparametrized conformal circle. Therefore, we can utilize the adapted coordinate of to avoid the tedious computation of the second fundamental forms. The following Lemma 4.3 and Proposition 4.4 respectively give the coordinate change of to its adapted coordinate and provide that satisfies (23) and (24), except for (25), for preserving proper surfaces.
Lemma 4.3.
Let be a Poincaré-Einstein space in the normal form relative to and be a surface orthogonal to with being a non-null curve. Assume it has a parametrization from to where and both at . The existence of the adapted coordinate of implies there is a coordinate change , with . The coordinate change is in the following modular higher orders.
(29) |
where
The partial derivatives of and above are at . The term in is . Note that and .
Proposition 4.4.
Proof.
Let be a non-null conformal circle in a coordinate open set of with the initial conditions and at . Let be its extended proper surface defined by the adapted coordinate in (26). Since is still an unparametrized conformal circle from Proposition 3.4, we know is a conformal local diffeomorphism due to Theorem 2.3. Without loss of generality, we assume that is the identity map on where and that the local coordinate of its extended local diffeomorphism is where is the geodesic defining function of . From Proposition 4.2 and Lemma 4.3, we let be the coordinate change of to its adapted coordinate.
Let , with and where is arbitrary near . The variable gives an -family of non-null conformal circles . The extended proper surfaces is defined by the formula of the adapted coordinate (26) where the coefficients in the expansion of depend on . Because depends smoothly on , we know the adapted coordinate for defined from Lemma 4.3 depends smoothly on . Since are proper surfaces, we have
(30) |
and
(31) |
Following the formulas and the conventions from (29), we know at the point . Using chain rule on for , the straightforward computation for (30) at is
(32) |
Therefore, and at . The results we got imply at and at from (29). Doing chain rule again, (31) is equal to the following at .
(33) |
So, and at .
Conversely, assume where we assume is the identity map on . Let be a proper surface and be its adapted coordinate. From Proposition 4.2, we know orthogonal to . Considering the coordinate change of to its adapted coordinate and following the formula from (29), we have
Computing and directly from chain rule, we have at
Theorem 4.5.
Let and be Poincaré-Einstein manifolds for and respectively with same signature . Given a local diffeomorphism such that it smoothly extends a local diffeomorphism . Assume satisfies
-
(i)
is non-null (resp. null) non-null (resp. null) if ,
-
(ii)
sgn= sgn if .
If the maps proper surfaces in to proper surfaces in , then there is a local diffeomorphism on an open neighborhood of ,
where smoothly extends the identity map on , such that is an asymptotic local isometry.
Proof.
From Proposition 4.4, we can choose and suitably to let be a local isometry. Consider the identification of in (18)
We aim to find out an open set such that it contains and the following map is well-defined
(34) |
where and .
For any , it has open neighborhoods in such that
where means its closure is compact in . Here we choose small enough such that the polynomial of on is strictly increasing for all . Hence, there exists so that
Let . Since we know the asymptotic expansion of from Proposition 4.4, we get is an asymptotic local isometry. ∎
Corollary 4.6.
Let and be local diffeomorphisms as stated in Theorem 4.5. Assume there is a geodesic defining function for some and such that
(35) |
where is the gradient of with respect to and is the geodesic defining function of to make be a local isometry.
-
(i)
If , then can be chosen as an embedding with its image while is small enough.
-
(ii)
If , then can be chosen to be the identity map on . Particularly, is an asymptotic local isometry.
Proof.
Recall the definition of in (34),
Let be small enough so that for some . If is small enough, then is strictly increasing because for
Hence, is an open injective immersion for . The case for is straightforward due to the definition of . ∎
The following proposition gives geometric conditions to satisfy the presumptions of Corollary 4.6.
Proposition 4.7.
Let and be local diffeomorphisms as stated in Theorem 4.5. Assume there is a geodesic defining function for some and such that
(36) |
Let be a geodesic defining function of to make be a local isometry.
-
(i)
If , there exists such that on .
-
(ii)
If , we have on .
Proof.
Consider the identification of near and in (18), . Then, (36) is equivalent to
where the right-hand side above is exactly from (21) while considering local conditions of an asymptotic local isometry. Then, the Taylor expansion for the right-hand side gives the following.
where , and mean the th-order partial derivative of at of , and respectively. So, we have from above
(37) |
where the third equality above is when at . We know , and at from Proposition 4.4. This completes the proof. ∎
References
- [1] T. N. Bailey and M. G. Eastwood. Conformal circles and parametrizations of curves in conformal manifolds. American Mathematical Society, 108(1):215–221, Jan. 1990.
- [2] T. N. Bailey, M. G. Eastwood, and A. R. Gover. Thomas’s structure bundle for conformal, projective and related structures. The Rocky Mountain Journal of Mathematics, 24(4):1191–1217, 1994.
- [3] Constantin Carathéodory. The most general transformations of plane regions which transform circles into circles. Bulletin of the American Mathematical Society, 43:573–579, 1937.
- [4] Constantin Carathéodory. The most general transformations of plane regions which transform circles into circles. Bulletin of the American Mathematical Society, 43:573–579, 1937.
- [5] Jih-Hsin Cheng. Chain-preserving diffeomorphisms and cr equivalence. American Mathematical Society, 103(1):75–80, 1988.
- [6] Sean Curry and A. Rod Gover. An introduction to conformal geometry and tractor calculus, with a view to applications in general relativity, 2014.
- [7] Boris Doubrov and Vojtěch Žádník. Equations and symmetries of generalized geodesics. 2005.
- [8] Charles Fefferman and C. Robin Graham. The Ambient Metric (AM-178). Annals of Mathematics Studies. Princeton University Press, 2011.
- [9] J. Fine and Y. Herfray. An ambient approach to conformal geodesics. Communications in Contemporary Mathematics, 24:2150009, 2022.
- [10] H. Friedrich and B. G. Schmidt. Conformal geodesics in general relativity. Royal Society, 414(1846):171–195, Nov. 9, 1987.
- [11] Helmut Friedrich. Conformal geodesics on vacuum space-times. Communications in Mathematical Physics, 235:513–543, 2003-04-01.
- [12] Jason Jeffers. Lost theorems of geometry. The American Mathematical Monthly, 107(9):800–812, 2000.
- [13] S. Kobayashi. A theorem on the affine transformation group of a riemannian manifold. Nagoya Mathematical Journal, 9:39–41, 1955.
- [14] S. Kobayashi and K. Nomizu. Foundations of Differential Geometry, volume 1. Interscience Publishers, 1963.
- [15] Christian Lübbe. A conformal extension theorem based on null conformal geodesics. Journal of Mathematical Physics, 50:112502, 2009.
- [16] Christian Lübbe and Paul Tod. An extension theorem for conformal gauge singularities. Journal of Mathematical Physics, 50(112501), 2009.
- [17] B O’Neill. Semi-Riemannian Geometry: With Applications to Relativity. Pure and applied mathematics. Academic Press, 1983.
- [18] R. Penrose and W. Rindler. Spinors and Space-Time: Volume 1, Two-Spinor Calculus and Relativistic Fields. Cambridge Monographs on Mathematical Physics. Cambridge University Press, 1984.
- [19] R. W. Sharpe. Differential Geometry: Cartan’s Generalization of Klein’s Erlangen Program. Springer, 1997.
- [20] Michael Spivak. A Comprehensive Introduction to Differential Geometry, volume II. Publish or Perish, Inc., 3rd edition, 1999.
- [21] Paul Tod. Some examples of the behaviour of conformal geodesics. Journal of Geometry and Physics, 62(8):1778–1792, 2012.
- [22] Edward Witten. Anti-de sitter space and holography. Advances in Theoretical and Mathematical Physics, 2:253–291, 1998.
- [23] Kentaro Yano. The Theory of Lie Derivatives and Its Applications. North-Holland Publishing Company, 1957.
- [24] Kentaro Yano and Yasuro Tomonaga. Quelques remarques sur les groupes de transformations dans les espaces à connexion linéaires, iv. Proceedings of the Japan Academy, 22:173–183, 1946.
- [25] Andreas Čap and A. Rod Gover. Standard tractors and the conformal ambient metric construction. Annals of Global Analysis and Geometry, 24:231–259, 2003-10-01.
- [26] Andreas Čap and Jan Slovák. Parabolic Geometries I: Background and General Theory. American Mathematical Society, 2009.