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

Constant angle null hypersurfaces

Samuel Chable-Naal, Matias Navarro, Didier A. Solis
Abstract

In this work we introduce the notion of constant angle null hypersurface of a Lorentzian manifold with respect to a given ambient vector field. We analyze the case in which the vector field is closed and conformal, thus finding that such null hypersurfaces have a canonical principal direction. We further provide some classification results for constant angle null surfaces with vanishing null mean curvature.

Keywords: Null hypersurfaces, closed conformal vector fields, constant angle hypersurfaces

MSC 2020 Classification: 53B30, 53A05, 53A10

1 Introduction

The classification of submanifolds subject to certain geometric restrictions has been one of the main driving forces in semi-Riemannian geometry. The characterization of totally geodesic, totally umbilical, minimal or isoparametric submanifolds of specific classes of ambient spaces are active areas of research up to this day (for detailed accounts refer to [7, 9, 12]). Historically, among the first such objects to be explored are the so-called constant angle submanifolds, its study dating back to the works of J. Bernoulli on spirals and rhumb lines [38].

In the Riemannian case, given an oriented immersed hypersurface MM of a Riemannian manifold (M¯,g¯)(\bar{M},\bar{g}) and a distinguished vector field VΓ(TM¯)V\in\Gamma(T\bar{M}), V0V\neq 0, the angle θ\theta of MM with respect to VV is defined by

cosθ=g¯(V,𝐧)|V|,\cos\theta=\frac{\bar{g}(V,\mathbf{n})}{|V|}, (1)

where 𝐧\mathbf{n} is a unit normal vector field to MM. Thus, a constant angle hypersurface is characterized by the constancy of θ\theta. This notion can be straightforwardly generalized to the semi-Riemannian context by requiring the right-hand size in Eq. (1) to be constant along MM. Notice that this latter expression carries a distinct geometrical meaning in some specific settings. For instance, in the Lorentzian case, it can be interpreted in terms of a hyperbolic cosine, when MM is spacelike and VV is timelike [35]. From the geometrical point of view, the most meaningful properties arise in the cases in which the vector field VV carries some relevant geometrical information. Classical examples of such fields include those that arise as infinitesimal generators of geometrical transformations like isometries (Killing fields) or conformal maps (closed conformal or concircular fields). It is worthwhile mentioning that the latter class include parallel, radial and gradient vector fields [27].

Several classification results for constant angle semi-Riemannian (non-degenerate) hypersurfaces have been established recently in ambient spaces such as cartesian products [10, 13, 16, 19], warped products [15, 22], spaceforms [20, 28, 30, 34] and other geometrically relevant spaces [1, 36, 37]. One remarkable relation between the distinguished vector field VV and the geometry of a constant angle hypersurface MM is encapsulated in the notion of canonical principal direction: the tangent component VΓ(TM)V^{\top}\in\Gamma(TM) is a principal direction of MM [11, 21, 23, 29].

We notice that the above discussion deals exclusively with non-degenerate hypersurfaces. Nevertheless, some of the most studied submanifolds in Lorentzian geometry and mathematical relativity —such as event and Killing horizons— are null, that is, degenerate [26, 25, 40]. In spite of its relevance in theoretical physics, a framework for the study of null submanifolds in a spirit close to the classical Riemannian submanifold theory was missing until the 1980’s. The main difficulty to be surmounted in this effort lies in the fact that in the realm of Lorentzian geometry the vector fields orthogonal to a null hypersurface MM¯M\subset\bar{M} are also tangent to it. Hence, there is no canonical orthogonal splitting of TM¯T\bar{M} into normal and tangent components. In [17] K. Duggal and A. Bejancu established the foundations for a null submanifold theory based on the choice of an additional structure, the so called screen distribution S(TM)S(TM). A considerable amount of research has developed following this approach (see for instance [18] and references therein). In particular, the study of null hypersurfaces in generalized Robertson-Walker spacetimes is an active area of research to this day [31].

2 Preliminaries

Following [17, 18], we denote by (M¯,g¯)(\bar{M},\bar{g}) an (n+2)(n+2)-dimensional Lorentzian manifold with metric g¯\bar{g} and by MM an (n+1)(n+1)-manifold immersed in M¯\bar{M} with degenerate induced metric gg. Equivalently, there exists a vector field ξ0\xi\neq 0 tangent to MM such that

g(ξ,X)=0,forallXΓ(TM).g(\xi,X)=0,\quad\mathrm{for\ all\ }X\in\Gamma(TM).

Since dimTpM+dimTpM=dimTpM¯\dim T_{p}M+\dim T_{p}M^{\perp}=\dim T_{p}\bar{M}, the radical space rad(TpM)=TpMTpM\mathrm{rad}(T_{p}M)=T_{p}M\cap T_{p}M^{\perp} at each pMp\in M satisfies

rad(TpM)=span(ξp).\mathrm{rad}(T_{p}M)=span(\xi_{p}).

The radical bundle rad(TM)\mathrm{rad}(TM) is given by

rad(TM)=pMrad(TpM).\mathrm{rad}(TM)=\bigcup\limits_{p\in M}\mathrm{rad}(T_{p}M).

A screen distribution on MM is a vector sub-bundle S(TM)S(TM) of TMTM such that

TM=S(TM)orthrad(TM),TM=S(TM)\oplus_{\mathrm{orth}}\mathrm{rad}(TM), (2)

thus, S(TM)S(TM) is complementary to rad(TM)\mathrm{rad}(TM) in TMTM.

Notice that due to the Lorentzian character of g¯\bar{g}, the metric restricted to S(TM)S(TM) is positive definite. Intuitively, an adequate choice of S(TM)S(TM) would enable us to relate the geometry of (M,g)(M,g) to a more familiar (Riemannian) geometry in S(TM)S(TM). Henceforth, by a null hypersurface we mean a triple (M,g,S(TM))(M,g,S(TM)).

The last main ingredient in this framework is the so called transversal bundle tr(TM)\mathrm{tr}(TM), which is the unique rank 11 vector bundle with the following property: given ξΓ(rad(TM))\xi\in\Gamma(\mathrm{rad}(TM)) there exists a unique NΓ(tr(TM))N\in\Gamma(\mathrm{tr}(TM)) such that

g¯(ξ,N)=1,g¯(N,N)=g¯(N,X)=0,forallXΓ(S(TM)).\bar{g}(\xi,N)=1,\quad\bar{g}(N,N)=\bar{g}(N,X)=0,\quad\mathrm{for\ all\ }X\in\Gamma(S(TM)). (3)

Thus, we can split the tangent bundle of the ambient manifold as

TM¯=TMtr(TM).T\bar{M}=TM\oplus\mathrm{tr}(TM). (4)
Remark 2.1.

Notice that even though a choice of screen distribution S(TM)S(TM) determines the transversal bundle tr(M)\mathrm{tr}(M), there is an scaling gauge for the sections ξ\xi and NN satisfying (3). Indeed, if the smooth function f:Mf:M\to\mathbb{R} never vanishes, then the smooth sections

ξ=fξ,N=1fN\xi^{\prime}=f\xi,\quad N^{\prime}=\frac{1}{f}N (5)

also satisfy (3).

Remark 2.2.

Let us consider a vector field ζ\zeta transversal to MM and ξΓ(rad(TM))\xi\in\Gamma(\mathrm{rad}(TM)). Then the vector field NζN_{\zeta} given by

Nζ=1g¯(ξ,ζ)(ζg¯(ζ,ζ)2g¯(ξ,ζ)ξ)N_{\zeta}=\frac{1}{\bar{g}(\xi,\zeta)}\left(\zeta-\frac{\bar{g}(\zeta,\zeta)}{2\bar{g}(\xi,\zeta)}\xi\right)

satisfies g¯(Nζ,Nζ)=0\bar{g}(N_{\zeta},N_{\zeta})=0 and g¯(Nζ,ξ)=1\bar{g}(N_{\zeta},\xi)=1. Therefore, ζ\zeta induces a screen distribution Sζ(TM)=span(ξ,Nζ)S_{\zeta}(TM)=span(\xi,N_{\zeta})^{\perp} (refer to Eq. 2.1.7 in [18])111Note that in this case ζΓ(Sζ(TM))\zeta\in\Gamma(S_{\zeta}(TM)^{\perp}).. The above construction is key in the rigging approach to null hypersurface theory [24].

In virtue of decompositions (2) and (4) we can introduce the notions of induced connections, shape operators and second fundamental forms, as well as the fundamental equations relating them.

Let ¯\bar{\nabla} be the Levi-Civita connection of (M¯,g¯)(\bar{M},\bar{g}) and VΓ(tr(TM))V\in\Gamma(\mathrm{tr}(TM)). Given X,YΓ(TM)X,Y\in\Gamma(TM), we can write the Gauss and Weingarten formulae related to decomposition (4) as

¯XY\displaystyle\bar{\nabla}_{X}Y =XY+h(X,Y),\displaystyle=\nabla_{X}Y+h(X,Y), (6)
¯XV\displaystyle\bar{\nabla}_{X}V =AVX+XtV,\displaystyle=-A_{V}X+\nabla_{X}^{t}V,

where XY,AVXΓ(TM)\nabla_{X}Y,A_{V}X\in\Gamma(TM), and h(X,Y),XtVΓ(tr(TM))h(X,Y),\nabla_{X}^{t}V\in\Gamma(\mathrm{tr}(TM)). The above induces linear connections \nabla and t\nabla^{t} on TMTM and tr(TM)\mathrm{tr}(TM), respectively. Notice however that \nabla is not a metric connection, though it is torsion-free. Further, AV:Γ(TM)Γ(TM)A_{V}:\Gamma(TM)\to\Gamma(TM) and hh is a symmetric section of Hom2(TM,TM;tr(TM))\mathrm{Hom}^{2}(TM,TM;\mathrm{tr}(TM)) that satisfies

(Xg)(Y,Z)=g¯(h(X,Y),Z)+g¯(Y,h(X,Z)).(\nabla_{X}g)(Y,Z)=\bar{g}(h(X,Y),Z)+\bar{g}(Y,h(X,Z)). (7)

for all X,Y,ZΓ(TM)X,Y,Z\in\Gamma(TM). \nabla, AVA_{V} and hh are called the induced connection, shape operator and second fundamental form of MM, respectively.

On the other hand, relative to decomposition (2) we have another set of Gauss-Weingarten formulae for TMTM:

XPY\displaystyle\nabla_{X}PY =XPY+h(X,PY),\displaystyle=\nabla_{X}^{*}PY+h^{*}(X,PY), (8)
XU\displaystyle\nabla_{X}U =AUX+XtU,\displaystyle=-A_{U}^{*}X+\nabla_{X}^{*t}U,

where UΓ(rad(TM))U\in\Gamma(\mathrm{rad}(TM)), PP denotes the projection of TMTM onto S(TM)S(TM); \nabla^{*} and t\nabla^{*t} are connections on S(TM)S(TM) and rad(TM)\mathrm{rad}(TM), respectively. Here AU:Γ(TM)Γ(S(TM))A_{U}^{*}:\Gamma(TM)\to\Gamma(S(TM)) and hHom2(TM,S(TM),rad(TM))h^{*}\in\mathrm{Hom}^{2}(TM,S(TM),\mathrm{rad}(TM)) are the screen shape operator and screen second fundamental form of S(TM)S(TM), respectively. A straightforward computation shows that \nabla^{*} is a metric connection with respect to g¯|S(TM)\bar{g}|_{S(TM)}.

For further reference, we cite some of the standard results pertaining null hypersurfaces that will be used throughout this work. For complete proofs, refer to Proposition 2.1.2 and 2.2.6 in [17] and [18], respectively.

Proposition 2.3.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface, UΓ(rad(TM))U\in\Gamma(\mathrm{rad}(TM)) and VΓ(tr(TM))V\in\Gamma(\mathrm{tr}(TM)). Then

  1. 1.

    AVA_{V} is S(TM)S(TM) valued.

  2. 2.

    AUU=0A^{*}_{U}U=0.

  3. 3.

    AξA^{*}_{\xi} is symmetric with respect to gg.

  4. 4.

    The following are equivalent:

    • S(TM)S(TM) is an integrable distribution.

    • hh^{*} is symmetric on S(TM)S(TM).

    • AVA_{V} is symmetric respect to g|S(TM)g|_{S(TM)}.

For YΓ(TM¯)Y\in\Gamma(T\bar{M}), applying decompositions (2) and (4) leads to

Y=Y+Yξ+YNY=\overset{*}{Y}+Y_{\xi}+Y_{N} (9)

where

Yξ=g¯(Y,N)ξ,YN=g¯(Y,ξ)NY_{\xi}=\overline{g}(Y,N)\xi,\quad Y_{N}=\overline{g}(Y,\xi)N

and Z\overset{*}{Z} is the projection of YY on S(TM)S(TM). We now establish the basic relations arising from coupling the above relation to the Gauss and Weingarten formulae when XΓ(S(TM))X\in\Gamma(S(TM)).

¯XY\displaystyle\bar{\nabla}_{X}Y =\displaystyle= ¯X(Y+Yξ)+¯XYN\displaystyle\bar{\nabla}_{X}(\overset{*}{Y}+Y_{\xi})+\bar{\nabla}_{X}Y_{N}
=\displaystyle= X(Y+Yξ)+h(X,Y+Yξ)AYNX+XtYN\displaystyle\nabla_{X}(\overset{*}{Y}+Y_{\xi})+h(X,\overset{*}{Y}+Y_{\xi})-A_{Y_{N}}X+\nabla_{X}^{t}Y_{N}
=\displaystyle= XY+XYξ+h(X,Y+Yξ)AYNX+XtYN\displaystyle\nabla_{X}\overset{*}{Y}+\nabla_{X}Y_{\xi}+h(X,\overset{*}{Y}+Y_{\xi})-A_{Y_{N}}X+\nabla_{X}^{t}Y_{N}
=\displaystyle= XY+h(X,Y)AYξX+XtYξ+h(X,Y)AYNX+XtYN.\displaystyle\overset{*}{\nabla}_{X}\overset{*}{Y}+\overset{*}{h}(X,\overset{*}{Y})-A_{Y_{\xi}}^{*}X+\nabla_{X}^{*t}Y_{\xi}+h(X,\overset{*}{Y})-A_{Y_{N}}X+\nabla_{X}^{t}Y_{N}.

In summary

¯XY=XYAYξXAYNXΓ(S(TM))+h(X,Y)+XtYξΓ(rad(TM))+h(X,Y)+XtYNΓ(tr(TM))\bar{\nabla}_{X}Y=\underbrace{\overset{*}{\nabla}_{X}\overset{*}{Y}-A_{Y_{\xi}}^{*}X-A_{Y_{N}}X}_{\in\Gamma(S(TM))}+\underbrace{\overset{*}{h}(X,\overset{*}{Y})+\nabla_{X}^{*t}Y_{\xi}}_{\in\Gamma(\mathrm{rad}(TM))}+\underbrace{h(X,\overset{*}{Y})+\nabla_{X}^{t}Y_{N}}_{\in\Gamma(\mathrm{tr}(TM))} (10)
Remark 2.4.

Notice that when S(TM)S(TM) is integrable their leaves form a foliation of MM by Riemannian manifolds (M^,g^)(\widehat{M},\hat{g}). As a consequence of Frobenius theorem, in such cases \overset{*}{\nabla} is torsion free and hence it coincides with the Levi-Civita connection ^\widehat{\nabla} of each leaf (M^,g^)(\widehat{M},\widehat{g}). Moreover, consider the Gauss-Weingarten formulae associated to decomposition TM¯=S(TM)S(TM)T\bar{M}=S(TM)\oplus S(TM)^{\perp}, that is,

¯XY=¯X(Y+Y)=^XY+α(X,Y)A^YX+XY,\bar{\nabla}_{X}Y=\bar{\nabla}_{X}(\overset{*}{Y}+Y^{\perp})=\widehat{\nabla}_{X}\overset{*}{Y}+\alpha(X,\overset{*}{Y})-\widehat{A}_{Y^{\perp}}X+\nabla_{X}^{\perp}Y^{\perp}, (11)

where α\alpha, A^Y\widehat{A}_{Y^{\perp}}, XY\nabla_{X}^{\perp}Y^{\perp} are the second fundamental form, shape operator and normal connections of M^\widehat{M} as a codimension two submanifold immersed in M¯\bar{M} (see [12]). Thus, comparing the S(TM)S(TM) components in Eqs. (10) and (11) leads to

A^YX=AYξX+AYNX,XΓ(S(TM)).\widehat{A}_{Y^{\perp}}X=A_{Y_{\xi}}X+A_{Y_{N}}X,\qquad X\in\Gamma(S(TM)). (12)

Let us now consider a null frame {ξ,N}\{\xi,N\} as in Eq. (3) and denote by BB, CC the bilinear forms associated to hh and hh^{*}, respectively. That is

h(X,Y)=B(X,Y)N,h(X,PY)=C(X,PY)ξ.h(X,Y)=B(X,Y)N,\qquad h^{*}(X,PY)=C(X,PY)\xi. (13)

Thus, in virtue of equations (6) and (8) we have

B(X,Y)\displaystyle B(X,Y) =g¯(h(X,Y),ξ)=g(Y,AξX),\displaystyle=\bar{g}(h(X,Y),\xi)=g(Y,A_{\xi}^{*}X), (14)
C(X,Y)\displaystyle C(X,Y) =g¯(h(X,PY),N)=g(PY,ANX),\displaystyle=\bar{g}(h^{*}(X,PY),N)=g(PY,A_{N}X),

for all X,YΓ(TM)X,Y\in\Gamma(TM). It then follows that

B(X,ξ)=0XΓ(TM).B(X,\xi)=0\quad\forall X\in\Gamma(TM). (15)

Notice that since hh is symmetric, then by Eq. (14) ANA_{N} is symmetric with respect to gg. Hence Proposition 2.3 yields that the same is true for AξA_{\xi} provided S(TM)S(TM) is integrable.

The null mean curvature HH is the trace of the shape operator AξA^{*}_{\xi}. Due to Eq. (15) we have

H=i=1nB(ei,ei),H=\sum_{i=1}^{n}B(e_{i},e_{i}),

where {e1,,en}\{e_{1},\ldots,e_{n}\} denotes an orthonormal frame field in S(TM)S(TM).

Furthermore, note that g¯(XtN,ξ)=g¯(Xtξ,N)\bar{g}(\nabla^{t}_{X}N,\xi)=-\bar{g}(\nabla_{X}^{*t}\xi,N). Hence, in terms of the one form τ\tau defined by

τ(X)=g¯(XtN,ξ)\tau(X)=\bar{g}(\nabla^{t}_{X}N,\xi) (16)

the Weingarten formulae read as

¯XV=AVX+τ(X)V,XU=AUXτ(X)U.\bar{\nabla}_{X}V=-A_{V}X+\tau(X)V,\quad\nabla_{X}U=-A_{U}^{*}X-\tau(X)U. (17)
Remark 2.5.

Notice that when τ0\tau\equiv 0 Eq. (17) looks just like the standard semi-Riemannian (non-degenerate) Weingarten formula for a unit normal vector field. In that sense, the vanishing of τ\tau can be interpreted as a normalization that ensures the closest possible analogue to the classical submanifold theory. As it turns out, a choice of {ξ,N}\{\xi,N\} that yields τ0\tau\equiv 0 is always possible if the Ricci tensor RicRic^{\nabla} vanishes (see Prop. 2.4.2 in [18]).

The curvature operators R¯\bar{R}, RR, RR^{*} associated to the linear connections ¯\bar{\nabla}, \nabla, \nabla^{*} are intertwined in a series of fundamental equations that resemble the standard Gauss-Codazzi structure equations in the non-degenerate case. We only present one of them, that will be used in Theorem 6.4 and refer to it as the Codazzi equation of MM. For all X,Y,ZΓ(TM)X,Y,Z\in\Gamma(TM) and UΓ(rad(TM))U\in\Gamma(\textrm{rad}(TM)) we have

g¯(R¯(X,Y)Z,U)=g¯((Xh)(Y,Z)(Yh)(X,Z),U)\bar{g}(\bar{R}(X,Y)Z,U)=\bar{g}((\nabla_{X}h)(Y,Z)-(\nabla_{Y}h)(X,Z),U)

Thus, once the null fields {ξ,N}\{\xi,N\} are fixed,

g¯(R¯(X,Y)Z,ξ)=(XB)(Y,Z)(YB)(X,Z)+τ(X)B(Y,Z)τ(Y)B(X,Z)\bar{g}(\bar{R}(X,Y)Z,\xi)=(\nabla_{X}B)(Y,Z)-(\nabla_{Y}B)(X,Z)+\tau(X)B(Y,Z)-\tau(Y)B(X,Z) (18)

where Eq. 7 translates to

(XB)(Y,Z)=XB(Y,Z)B(XY,Z)B(Y,XZ).(\nabla_{X}B)(Y,Z)=X\cdot B(Y,Z)-B(\nabla_{X}Y,Z)-B(Y,\nabla_{X}Z).

Many of the most significant geometrical restrictions widely studied in the semi-Riemannian (non degenerate) theory of submanifolds have analogues in the null setting. For instance, MM is totally umbilical (resp. totally geodesic) if there exists a smooth function λ\lambda such that h(X,Y)=λg(X,Y)h(X,Y)=\lambda g(X,Y) (resp. h0h\equiv 0). Similarly, S(TM)S(TM) is totally umbilical (resp. totally geodesic) if there exists a smooth function λ\lambda such that h(X,Y)=λg(X,Y)h^{*}(X,Y)=\lambda g(X,Y) (resp. h0h^{*}\equiv 0). In terms of the shape operators, we have that MM is totally umbilical (resp. totally geodesic) if AξPX=λPXA_{\xi}^{*}PX=\lambda PX (Aξ0A_{\xi}^{*}\equiv 0) and likewise, S(TM)S(TM) is totally umbilical (resp. totally geodesic) if ANX=λPXA_{N}X=\lambda PX (AN0A_{N}\equiv 0) for all XΓ(TM)X\in\Gamma(TM). Notice, that all these conditions, as well as the vanishing of the null mean curvature (H0H\equiv 0) do not depend on the choice of the null frame {ξ,N}\{\xi,N\}.

As an illustrative example, let us focus on the case of generalized Robertson-Walker spacetimes, or GRW spacetimes for short. Let us recall that they are defined as Lorentzian warped products of the form M¯=I×ϱF\bar{M}=-I\times_{\varrho}F, with (F,gF)(F,g_{F}) a Riemannian manifold and ϱ\varrho a positive smooth function defined on the real interval II. Thus we have

g¯=σ(dt2)+(ϱπ)2π(gF).\bar{g}=-\sigma^{*}(dt^{2})+(\varrho\circ\pi)^{2}\pi^{*}(g_{F}). (19)

where σ\sigma and π\pi are the projections of M¯\bar{M} over II and FF, respectively. If the fiber FF is a Riemannian manifold of constant sectional curvature suitable choices of the warping function deliver open sets of all Lorentzian spaceforms as follows:

  • Lorentz-Minkowski space: 𝕃n+2=×idn+1\mathbb{L}^{n+2}=-\mathbb{R}\times_{\text{id}}\mathbb{R}^{n+1}.

  • de Sitter space: 𝕊1n+2=×cosh𝕊n+1\mathbb{S}_{1}^{n+2}=-\mathbb{R}\times_{\cosh}\mathbb{S}^{n+1}.

  • Anti de Sitter space 1n+2=(π2,π2)×cosn+1\mathbb{H}_{1}^{n+2}=-\left(-\frac{\pi}{2},\frac{\pi}{2}\right)\times_{\cos}\mathbb{H}^{n+1}.

Example 2.6.

Let f:Ff:F\to\mathbb{R} be a transnormal function, that is, a smooth function such that |gradf|=ϱf|\text{grad}\,f|=\varrho\circ f. Thus the graph M={(f(p),p)|pF}M=\{\,(f(p),p)\,|\,p\in F\,\} is a degenerate hypersurface in the GRW spacetime M¯=I×ϱF\bar{M}=-I\times_{\varrho}F. Further, let S(TM)S^{*}(TM) be the family of tangent bundles of the level hypersurfaces St=M({t}×F)S_{t}=M\cap\left(\{t\}\times F\right). Consider now ξΓ(rad(TM))\xi\in\Gamma(\mathrm{rad}(TM)) and NΓ(tr(TM))N\in\Gamma(\mathrm{tr}(TM)) given by

ξ=12(t+1(ϱf)2gradf¯),N=12(t+1(ϱf)2gradf¯),\xi=\frac{1}{\sqrt{2}}\left(\partial_{t}+\frac{1}{(\varrho\circ f)^{2}}\overline{\text{grad}f}\right),\quad N=\frac{1}{\sqrt{2}}\left(-\partial_{t}+\frac{1}{(\varrho\circ f)^{2}}\overline{\text{grad}f}\right),

where gradf¯\overline{\text{grad}f} denotes the lift of gradf\text{grad}f to Γ(TM¯)\Gamma(T\bar{M}). In this setting we have that the shape operators of (M¯,g,S(TM))(\bar{M},g,S^{*}(TM)) satisfy

12(ANAξ)=ϱϱP.\frac{1}{\sqrt{2}}(A_{N}-A_{\xi}^{*})=\frac{\varrho^{\prime}}{\varrho}P. (20)

The above example serves as motivation for the concept of screen quasi-conformal hypersurface, which will be studied in the next section. For a detailed account, see [33, 32].

2.1 Closed conformal vector fields

A property that characterizes GRW spacetimes relates to the notion of closed conformal vector field. Let us recall that ZΓ(TM¯)Z\in\Gamma(T\bar{M}) is closed conformal —or CC for short— if there exists a smooth function φ:M¯\varphi:\bar{M}\to\mathbb{R} such that

¯XZ=φX\bar{\nabla}_{X}{Z}=\varphi X (21)

for all XΓ(TM¯)X\in\Gamma(T\bar{M}). Indeed, GRW spacetimes can also be described as those Lorentzian manifolds having a timelike CC vector field [8]. Throughout this work, φ\varphi will always denote the function associated to a CC vector field.

In the case of semi-Riemannian spaceforms, when viewed as hyperquadrics immersed in a semi-Euclidean space, CC vector fields arise as projections of parallel vector fields defined on the respective semi-Euclidean space [34, 21]. Notice that according to Eq. (19), the vector field Z=ϱtZ=\varrho\partial_{t} of Example 2.6 is closed conformal.

Other important classes of CC vector fields include parallel (φ0\varphi\equiv 0), homothetic (φ\varphi is a constant function) and radial, that is, when M¯\bar{M} is a semi-Euclidean space and Z(x)=φx+Z0Z(x)=\varphi x+Z_{0}, with Z0Z_{0} a parallel vector field.

One of the most remarkable features of CC vector fields is that their length |Z|=|g¯(Z,Z)||Z|=\sqrt{|\bar{g}(Z,Z)|} is constant (if not null) along directions orthogonal to them (see Lemma 2.7 in [34]).

Lemma 2.7.

Let ZΓ(TM¯)Z\in\Gamma(T\bar{M}) be a CC vector field and XΓ(TM¯)X\in\Gamma(T\bar{M}) such that g¯(Z,X)=0\bar{g}(Z,X)=0. If g¯(Z,Z)\bar{g}(Z,Z) never vanishes, then X|Z|=0X\cdot|Z|=0.

Proof.

Notice that

Xg¯(Z,Z)=2g¯(¯XZ,Z)=2φg¯(X,Z)=0.X\cdot\bar{g}(Z,Z)=2\bar{g}(\bar{\nabla}_{X}Z,Z)=2\varphi\bar{g}(X,Z)=0.

The result follows at once. ∎

In very recent times, we have witnessed an increasing interest in exploring the interplay between a CC vector field and the geometry of certain classes of null hypersurfaces when the ambient manifold is a spaceform [39], a space of quasi constant curvature [5], or in relation to the Raychaudhuri equation [4]. A common starting point consists in describing a CC vector field ZZ in terms of decomposition (9), that is, Z=Z+Zξ+ZNZ=\overset{*}{Z}+Z_{\xi}+Z_{N}, where Zξ=g¯(Z,N)ξZ_{\xi}=\overline{g}(Z,N)\xi, ZN=g¯(Z,ξ)NZ_{N}=\overline{g}(Z,\xi)N and ZΓ(S(TM))\overset{*}{Z}\in\Gamma(S(TM)). We now establish the basic relations arising from coupling the above relation to the Gauss and Weingarten formulae. Compare the following result with Proposition 3.1 in [5] and Proposition 3.7 in [39].

Lemma 2.8.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}). If ZΓ(TM¯)Z\in\Gamma(T\bar{M}) is a CC vector field, then for any XΓ(S(TM))X\in\Gamma(S(TM)):

  1. (a)

    φX=XZg¯(Z,N)AξXg(Z,ξ)ANX,\varphi X=\overset{*}{\nabla}_{X}\overset{*}{Z}-\overline{g}(Z,N)A^{*}_{\xi}X-g(Z,\xi)A_{N}X,

  2. (b)

    C(X,Z)+Xg(Z,N)g(Z,N)τ(X)=0C(X,\overset{*}{Z})+X\cdot g(Z,N)-g(Z,N)\tau(X)=0 and

  3. (c)

    B(X,Z)+Xg(Z,ξ)+g(Z,ξ)τ(X)=0B(X,\overset{*}{Z})+X\cdot g(Z,\xi)+g(Z,\xi)\tau(X)=0.

Proof.

By Eq. 10 we have

φX=(XZAZξXAZNX)+(h(X,Z+XtZξ)+(h(X,Z)+XtZN)\displaystyle\varphi X=(\overset{*}{\nabla}_{X}\overset{*}{Z}-A_{Z_{\xi}}^{*}X-A_{Z_{N}}X)+(\overset{*}{h}(X,\overset{*}{Z}+\nabla_{X}^{*t}Z_{\xi})+(h(X,\overset{*}{Z})+\nabla_{X}^{t}Z_{N})

Thus comparing the S(TM)S(TM) components in both sides of the above equation yields

φX=XZg(Z,N)AξXg(Z,ξ)ANX\varphi X=\overset{*}{\nabla}_{X}\overset{*}{Z}-g(Z,N)A_{\xi}^{*}X-g(Z,\xi)A_{N}X

so (a) holds. Similarly, by looking at the rad(TM)\mathrm{rad}(TM) and tr(M)\mathrm{tr}(M) components we get

0\displaystyle 0 =\displaystyle= h(X,Z)+XtZN,\displaystyle h(X,\overset{*}{Z})+\nabla_{X}^{t}Z_{N},
0\displaystyle 0 =\displaystyle= h(X,Z)+XtZξ.\displaystyle\overset{*}{h}(X,\overset{*}{Z})+\overset{*}{\nabla}_{X}^{t}Z_{\xi}.

Recall that h(X,Z)=B(X,Z)Nh(X,\overset{*}{Z})=B(X,\overset{*}{Z})N and thus

XtZN=Xtg(Z,ξ)N=((Xg(Z,ξ))+g(Z,ξ)τ(X))N.\nabla_{X}^{t}Z_{N}=\nabla_{X}^{t}g(Z,\xi)N=((X\cdot g(Z,\xi))+g(Z,\xi)\tau(X))N.

so (c) holds as well. Finally, (b) follows from an argument analogous to the proof of (c) above. ∎

As an immediate consequence of Lemma 2.8 (a) we can describe the cases in which ZZ lies entirely on rad(TM)\mathrm{rad}(TM) or tr(TM)\mathrm{tr}(TM) (compare to Thrm. 3.9 in [39]).

Corollary 2.9.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field. Then

  1. 1.

    If Z=Zξ=0\overset{*}{Z}=Z_{\xi}=0 , then MM is totally umbilical.

  2. 2.

    If Z=ZN=0\overset{*}{Z}=Z_{N}=0, then S(TM)S(TM) is totally umbilical.

Moreover, if ZZ is parallel then

  1. 1.

    If Z=Zξ=0\overset{*}{Z}=Z_{\xi}=0 , then MM is totally geodesic.

  2. 2.

    If Z=ZN=0\overset{*}{Z}=Z_{N}=0, then S(TM)S(TM) is totally geodesic.

The following Lemma will be key in several results in Sections 4 y 5.

Lemma 2.10.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field. If S(TM)S(TM) in integrable, then

X(g¯(X,ξ)g¯(X,N))=g¯(X,AZZ),XΓ(S(TM)).X\cdot(\bar{g}(X,\xi)\bar{g}(X,N))=-\bar{g}(X,A_{Z^{\perp}}\overset{*}{Z}),\quad X\in\Gamma(S(TM)).
Proof.

Since S(TM)S(TM) is integrable then ANA_{N} is symmetric. Thus by points (b) and (c) of Lemma 2.8 and Eq (12) we have

X(g¯(X,ξ)g¯(X,N))\displaystyle X\cdot(\bar{g}(X,\xi)\bar{g}(X,N)) =g¯(Z,ξ)g¯(X,ANZ)g¯(Z,N)g¯(X,AξZ)\displaystyle=-\bar{g}(Z,\xi)\bar{g}(X,A_{N}\overset{*}{Z})-\bar{g}(Z,N)\bar{g}(X,A^{*}_{\xi}\overset{*}{Z})
=g¯(X,AZNZ)g¯(X,AZξZ)\displaystyle=-\bar{g}(X,A_{Z_{N}}\overset{*}{Z})-\bar{g}(X,A^{*}_{Z_{\xi}}\overset{*}{Z})
=g¯(X,AZZ).\displaystyle=-\bar{g}(X,A_{Z^{\perp}}\overset{*}{Z}).

We end up this section with the computation of the screen gradient of the norm of a CC vector field ZZ.

Lemma 2.11.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field. Then

|Z|=εZφ|Z|Z,\overset{*}{\nabla}|Z|=\frac{\varepsilon_{Z}\varphi}{|Z|}\overset{*}{Z},

where εZ=±1\varepsilon_{Z}=\pm 1 is the sign of g¯(Z,Z)\overline{g}(Z,Z).

Proof.

Given an orthonormal frame {e1,e2,,en}\{e_{1},e_{2},\ldots,e_{n}\} of S(TM)S(TM) we have

g¯(Z,Z)\displaystyle\overset{*}{\nabla}\bar{g}(Z,Z) =i=1n(eig¯(Z,Z))ei=i=1n2g¯(¯eiZ,Z)ei=i=1n2g¯(φei,Z)ei\displaystyle=\sum_{i=1}^{n}(e_{i}\cdot\bar{g}(Z,Z))e_{i}=\sum_{i=1}^{n}2\bar{g}(\overline{\nabla}_{e_{i}}Z,Z)e_{i}=\sum_{i=1}^{n}2\bar{g}(\varphi e_{i},Z)e_{i}
=2φZ.\displaystyle=2\varphi\overset{*}{Z}.

Finally

|Z|=εZg(Z,Z)2|Z|=εZφ|Z|Z.\overset{*}{\nabla}\lvert Z\rvert=\frac{\overset{*}{\nabla}\varepsilon_{Z}g(Z,Z)}{2\lvert Z\rvert}=\frac{\varepsilon_{Z}\varphi}{\lvert Z\rvert}\overset{*}{Z}.

3 Screen quasi-conformal null hypersurfaces

One key question that arises naturally is whether the geometry of S(TM)S(TM) as a subbundle of TMTM is related in a natural way to its geometry as a subbundle of TM¯T\bar{M}. In [3] K. Duggal and C. Atindogbe define the notion of screen conformal null hypersurface and show that the answer is affirmative for this class of hypersurfaces. In precise terms, a null hypersurface (M,g,S(TM))(M,g,S(TM)) is screen conformal if its shape operators are linearly dependent, so there exists ϕ𝒞(M)\phi\in\mathcal{C}^{\infty}(M) such that

AN=ϕAξ.A_{N}=\phi A_{\xi}^{*}.

Numerous classification results exist for such hypersurfaces when the ambient manifold (M¯,g¯)(\bar{M},\bar{g}) has constant curvature [18]. However, in virtue of Example 2.6 an important class of space-times does not fit in this context. Thus, in order to incorporate such examples, the notion of screen quasi-conformal null hypersurface was introduced in [32].

Definition 3.1.

A null hypersurface (M,g,S(TM))(M,g,S(TM)) of (M¯,g¯)(\bar{M},\bar{g}) is screen quasi-conformal if the shape operators ANA_{N} and AξA_{\xi}^{*} of MM and S(TM)S(TM) satisfy

AN=ϕAξ+ψP,A_{N}=\phi A_{\xi}^{*}+\psi P,

for some functions ϕ,ψ𝒞(M)\phi,\psi\in\mathcal{C}^{\infty}(M); where P:Γ(TM)Γ(S(TM))P:\Gamma(TM)\to\Gamma(S(TM)) is the natural projection. We call (ϕ,ψ)(\phi,\psi) a quasi-conformal pair.

Screen quasi-conformal null hypersurfaces satisfying classical geometric restrictions (totally geodesic, totally umbilical, isoparametric, Einstein) have been studied recently (see [31] and references therein). Moreover, if (M,g,S(TM))(M,g,S(TM)) is screen quasi-conformal, then S(TM)S(TM) is integrable (see Theorem 3.7 in [32]). As can be readily checked, the null hypersurfaces (M,g,S(TM))(M,g,S^{*}(TM)) depicted in Example 2.6 are screen quasi-conformal with quasi-conformal pair (1,2ϱ/ϱ)(1,\sqrt{2}\varrho^{\prime}/\varrho). 222Notice that in this case the following additional conditions hold: (a) ϕ\phi and ψ\psi are constant along the leaves of S(TM)S(TM), and (b) τ(X)=0\tau(X)=0 for all XΓ(S(TM))X\in\Gamma(S(TM)).

At this point it is natural to ask under which circumstances we can endow a degenerate hyperfurface with a screen distribution in a way that the corresponding null hyperfurface (M,g,S(TM))(M,g,S(TM)) is screen conformal. In [18], Theorem 2.3.5 establishes that if the ambient space admits a parallel vector field, then it induces a screen conformal structure in any null hyperfurface. As the next result shows, there is an analogous connection between closed and conformal vector fields and screen quasi-conformal distributions, thus generalizing the aforementioned result.

Theorem 3.2.

Let (M,g)(M,g) be a degenerate hypersurface immersed (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a closed and conformal vector field with g¯(Z,Z)0\bar{g}(Z,Z)\neq 0. If ZZ is nowhere tangent to MM then there exists a screen distribution S(TM)S^{\prime}(TM) so that (M,g,S(TM))(M,g,S^{\prime}(TM)) is screen quasi-conformal. Furthermore, τ(X)=0\tau(X)=0 XΓ(S(TM))\forall X\in\Gamma(S^{\prime}(TM)).

Proof.

Since ZpTpMZ_{p}\not\in T_{p}M we have that E=span(Z)Rad(TM)E=span(Z)\oplus\mathrm{Rad}(TM) is a rank 22 vector bundle in which g¯|E\bar{g}|_{E} has Lorentzian character. Consider S(TM)=ES^{\prime}(TM)=E^{\perp} and let tr(TM)\mathrm{tr}(TM) be a null line bundle complementary to Rad(TM)\mathrm{Rad}(TM) in EE. Since ZΓ(E)Z\in\Gamma(E) we have ZrΓ(Rad(TM))Z_{r}\in\Gamma(\mathrm{Rad}(TM)) and ZtΓ(tr(TM))Z_{t}\in\Gamma(\mathrm{tr}(TM)) such that

Z=Zr+ZtZ=Z_{r}+Z_{t}

Hence, we can choose ξΓ(Rad(TM))\xi\in\Gamma(\mathrm{Rad}(TM)) and NΓ(Rad(TM))N\in\Gamma(\mathrm{Rad}(TM)) as

ξ=Zr,N=1g¯(Zr,Zt)Zt,\xi=Z_{r},\quad N=\frac{1}{\bar{g}(Z_{r},Z_{t})}Z_{t},

or equivalently

Z=ξ+θN,Z=\xi+\theta N,

where

θ=g¯(Zr,Zt)=12g¯(Z,Z).\theta=\bar{g}(Z_{r},Z_{t})=\frac{1}{2}\bar{g}(Z,Z).

Thus, for XΓ(TM)X\in\Gamma(TM) we have

φX\displaystyle\varphi X =\displaystyle= ¯XZ\displaystyle\bar{\nabla}_{X}Z
=\displaystyle= ¯Xξ+(Xθ)N+θ¯XN\displaystyle\bar{\nabla}_{X}\xi+(X\cdot\theta)N+\theta\bar{\nabla}_{X}N
=\displaystyle= Xξ+h(X,ξ)+(Xθ)N+θ(ANX+τ(X)N)\displaystyle\nabla_{X}\xi+h(X,\xi)+(X\cdot\theta)N+\theta(-A_{N}X+\tau(X)N)
=\displaystyle= AξXθANXτ(X)ξ+(Xθ)N+θτ(X)N.\displaystyle-A^{*}_{\xi}X-\theta A_{N}X-\tau(X)\xi+(X\cdot\theta)N+\theta\tau(X)N.

Thus we immediately have

φPX\displaystyle\varphi PX =\displaystyle= AξXθANX,\displaystyle-A^{*}_{\xi}X-\theta A_{N}X,
0\displaystyle 0 =\displaystyle= (Xθ)N+θτ(X)N.\displaystyle(X\cdot\theta)N+\theta\tau(X)N.

Hence, the first equation implies that (M,g,S(TM))(M,g,S^{\prime}(TM)) is screen quasi-conformal. Moreover, by Lemma 2.7, θ0\theta\neq 0 is constant along S(TM)S^{\prime}(TM) and the second equation implies τ(X)=0\tau(X)=0. ∎

Finally, let us notice that if the CC field is orthogonal to the screen distribution, then (M,g,S(TM))(M,g,S(TM)) is screen quasi-conformal.

Proposition 3.3.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}). If ZΓ(TM¯)Z\in\Gamma(T\bar{M}) is a CC vector field such that Z=0\overset{*}{Z}=0, then (M,g,S(TM))(M,g,S(TM)) is screen quasi-conformal.

Proof.

By the same calculations as in Theorem 3.2 we get

φX\displaystyle\varphi X =g(Z,N)AξXg(Z,ξ)ANX,\displaystyle=-g(Z,N)A^{*}_{\xi}X-g(Z,\xi)A_{N}X,

so we obtain

ANX=g(Z,N)g(Z,ξ)AξX+φg(Z,ξ)X.\displaystyle A_{N}X=-\frac{g(Z,N)}{g(Z,\xi)}A^{*}_{\xi}X+\frac{\varphi}{g(Z,\xi)}X.

4 Constant angle null hypersurfaces

As it is often the case, dealing with degenerate submanifolds requires some adaptation of the standard definitions of classical geometrical concepts. For the case of a semi-Riemannian hypersurface, the notion of angle respect to a vector field VV, as given by Eq. (1), heavily depends upon the normalization of a vector field orthogonal to the hypersurface, thus providing a unit normal vector field 𝐧\mathbf{n} along it. For a null hypersurface (M,g,S(TM))(M,g,S(TM)) the absence of a canonical normalization in the null directions ξ\xi and NN rules out the possibility for the functions

(V,ξ):=g¯(V,ξ)|V|,(V,N):=g¯(V,N)|V|,\measuredangle(V,\xi):=\frac{\bar{g}(V,\xi)}{|V|},\quad\measuredangle(V,N):=\frac{\bar{g}(V,N)}{|V|},

to be suitable candidates for describing an angle between MM and a nowhere null vector field VV.

Indeed, the gauge freedom of Eq. (5) enables that such functions (when non vanishing) might take any possible value after a scaling. In particular, we can always choose ff such that (V,ξ^)\measuredangle(V,\hat{\xi}) is constant. For instance, by taking f=g¯(V,ξ)/|Z|f=\bar{g}(V,\xi)/|Z| we get (V,ξ^)=1\measuredangle(V,\hat{\xi})=1. A similar situation holds for (V,N^)\measuredangle(V,\hat{N}). Notice however, that in general we can not use Eq. (5) in order to make both functions (V,ξ^)\measuredangle(V,\hat{\xi}) and (V,N^)\measuredangle(V,\hat{N}) simultaneously constant. Thus, in the spirit of keeping the algebraic form of Eq. (1), we consider the following definition.

Definition 4.1.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and VV be a nowhere null vector field along MM. We say that (M,g,S(TM))(M,g,S(TM)) has constant angle respect to VV if there exists ξΓ(rad(TM))\xi\in\Gamma(\textrm{rad}(TM)) and NΓ(tr(TM))N\in\Gamma(\textrm{tr}(TM)) such that the functions

(V,ξ)=g¯(V,ξ)|V|and(V,N)=g¯(V,N)|V|\measuredangle(V,\xi)=\frac{\bar{g}(V,\xi)}{|V|}\quad\textrm{and}\quad\measuredangle(V,N)=\frac{\bar{g}(V,N)}{|V|} (22)

are constant.

At first sight, one could object that Definition 4.1 is not very useful in concrete examples, as the non-constancy of the angle functions (22) for a given pair of null vector fields {ξ,N}\{\xi,N\} is indecisive as whether (M,g,S(TM))(M,g,S(TM)) has constant angle respect to VV or not. However, as the next result shows, we can formulate Definition 4.1 in terms that do not depend on any particular choice of a scaling.

Lemma 4.2.

Let (M,g,S(TM))(M,g,S(TM)) be a null hypersurface in (M¯,g¯)(\bar{M},\bar{g}) and VV a nowhere null vector field along MM. Then the following statements are equivalent:

  1. 1.

    (M,g,S(TM))(M,g,S(TM)) has constant angle with respect to ZZ.

  2. 2.

    For any choice of ξΓ(rad(M))\xi\in\Gamma(\textrm{rad}(M)) and NΓ(tr(M))N\in\Gamma(\textrm{tr}(M)) the function (V,ξ)(V,N)\measuredangle(V,\xi)\measuredangle(V,N) is constant.

  3. 3.

    g(V/|V|,V/|V|)g(V^{*}/|V|,V^{*}/|V|) is constant.

Proof.

We first show the equivalence between conditions (1) and (2). Let us assume (V,ξ)\measuredangle(V,\xi) and (V,N)\measuredangle(V,N) are constant. Then for ξ^=fξ\hat{\xi}=f\xi and N^=N/f\hat{N}=N/f we have that (V,ξ^)(V,N^)=(V,ξ)(V,N)\measuredangle(V,\hat{\xi})\measuredangle(V,\hat{N})=\measuredangle(V,\xi)\measuredangle(V,N), so it is clearly a constant function.

Conversely, assume k=(V,ξ)(V,N)k=\measuredangle(V,\xi)\measuredangle(V,N) is constant. Take

f=(V,N)k,f=\frac{\measuredangle(V,N)}{k},

hence

(V,ξ^)=f(V,ξ)=1,(V,N^)=(V,N)/f=k.\measuredangle(V,\hat{\xi})=f\measuredangle(V,\xi)=1,\quad\measuredangle(V,\hat{N})=\measuredangle(V,N)/f=k.

To establish the equivalence between conditions (2) and (3) notice that

g(V,V)=g(V,V)+2g(V,ξ)g(V,N),g(V,V)=g(V^{*},V^{*})+2g(V,\xi)g(V,N), (23)

thus

g(V|V|,V|V|)=g(V,V)|V|22g(V,ξ)|V|g(V,N)|V|=εZ2(V,ξ)(V,N)g\left(\frac{V^{*}}{|V|},\frac{V^{*}}{|V|}\right)=\frac{g(V,V)}{|V|^{2}}-2\frac{g(V,\xi)}{|V|}\frac{g(V,N)}{|V|}=\varepsilon_{Z}-2\measuredangle(V,\xi)\measuredangle(V,N)

and the result follows immediately.

Recall that in the semi-Riemannian setting, constant angle hypersurfaces can be defined via the squared norm of V/|V|V^{\top}/|V|, or equivalently, the squared norm of V/|V|V^{\perp}/|V|; where VV^{\top} and VV^{\perp} denote the tangent and normal components of VV. Thus Lemma 4.2 can be interpreted as a generalization to the null context of this classical result (see for instance Lemma 3.3 in [34]).

Furthermore, two exceptional cases arise naturally when dealing with non degenerate hypersurfaces (M,g)(M,g), namely, when the vector field VV is orthogonal to the hypersurface and when it is tangent to it. For instance, if the normal distribution to VV is integrable, then a unit normal vector field to one of its integral manifolds in necessarily collinear to VV, and thus we can think of the hypersurface as making a zero angle with respect to VV. This is for instance the case of hyperplanes or hyperquadrics in semi-Euclidean spaces when we consider parallel or radial vector fields333Recall that CC vector fields in semi-Euclidean spaces are necesarily either parallel or radial.. On the other hand, if VV is tangent to the hypersurface (M,g)(M,g) then VV is orthogonal to any unit normal vector field, as it is indeed the case for cylinders (VV parallel) and cones (VV radial) in semi-Euclidean spaces.

Example 4.3.

In view of Lemma 4.2 we have that the proper analogue for a null hypersurface (M,g,S(TM))(M,g,S(TM)) making a zero angle with respect to a vector field VV are precisely those for which V=0\overset{*}{V}=0, that is, when VΓ(S(TM))V\in\Gamma(S(TM)^{\perp}). Notice that if we assume VV to be a CC vector field, then Proposition 2.9 ensures that totally geodesic or totally umbilic null hypersurfaces fit in this description, in close analogy to the non degenerate case. Moreover, if we take a vector field ζΓ(TM¯)\zeta\in\Gamma(T\bar{M}) transversal to MM and consider the associated screen distribution Sζ(TM)S_{\zeta}(TM) as in Remark 2.2, then ζ=0\overset{*}{\zeta}=0 and consequently (M,g,Sζ(TM))(M,g,S_{\zeta}(TM)) makes a constant angle with respect to ζ\zeta.

Example 4.4.

Let us consider a null hyperplane Π1n+2\Pi\subset\mathbb{R}^{n+2}_{1} and a parallel vector field VΓ(T1n+2)V\in\Gamma(T\mathbb{R}^{n+2}_{1}). We can always choose ξ\xi to be parallel along Π\Pi, and thus (V,ξ)\measuredangle(V,\xi) is constant. Moreover, if we take S(TΠ)S^{\prime}(T\Pi) as the tangent spaces to a family of parallel planar sections of Π\Pi then NN is also parallel along MM and thus (V,N)\measuredangle(V,N) is constant as well. Consequently, (Π,g,S(TΠ))(\Pi,g,S^{\prime}(T\Pi)) is a constant angle null hypersurface of 1n+2\mathbb{R}^{n+2}_{1}.

Conversely, assume (Π,g,S(TΠ))(\Pi,g,S(T\Pi)) is a constant angle null hypersurface with respect to a parallel vector field VV. Further, without loss of generality we can assume V=e0V=e_{0} if VV is timelike, choose ξΓ(rad(TM))\xi\in\Gamma(\textrm{rad}(TM)) such that (V,ξ)=g¯(V,ξ)=1\measuredangle(V,\xi)=\bar{g}(V,\xi)=1. Let N=(N0,N1,,Nn+1)N=(N_{0},N_{1},\ldots,N_{n+1}) such that (N,V)=N0\measuredangle(N,V)=-N_{0} is constant. Thus N22++Nn+12=1N_{2}^{2}+\cdots+N_{n+1}^{2}=1 and

S(TM)=span{e2N2ξ,,en+1Nn+1ξ}.S(TM)=span\{e_{2}-N_{2}\xi,\ldots,e_{n+1}-N_{n+1}\xi\}.

Thus, for n=1n=1 we have that (Π,g,S(TM))(\Pi,g,S(TM)) is constant angle hypersurface only if S(TM)S(TM) is planar, that is, S(TM)pS(TM)_{p} is generated by the intersection of TpMT_{p}M and a hyperplane in the ambient space 1m+2\mathbb{R}^{m+2}_{1} passing through pp (see Figure 1). On the other hand, for n2n\geq 2 we have infinitely many non planar null hypersurfaces.

Refer to caption
Figure 1: A planar constant angle null hypersurface.
Example 4.5.

We now provide some examples of constant angle null hypersurfaces in the two-dimensional light cone Λ12\Lambda^{2}_{1}. Let ZZ be a timelike parallel vector field along Λ1213\Lambda^{2}_{1}\subset\mathbb{R}^{3}_{1}. Without loss of generality, we can assume Z=e0Z=e_{0}. Let Φ:2Λ12\Phi:\mathbb{R}^{2}\rightarrow\Lambda^{2}_{1} given by Φ(a1,a2)=(a12+a22,a1,a2)\Phi(a_{1},a_{2})=(\sqrt{a_{1}^{2}+a_{2}^{2}},a_{1},a_{2}) be a graph parametrization of the light cone, then

ξ=Φ/a12+a222e0=(1,a1a12+a22,a2a12+a22)\xi=\Phi/\sqrt{a_{1}^{2}+a_{2}^{2}}-2e_{0}=\left(-1,\dfrac{a_{1}}{\sqrt{a_{1}^{2}+a_{2}^{2}}},\dfrac{a_{2}}{\sqrt{a_{1}^{2}+a_{2}^{2}}}\right)

satisfies g(Z,ξ)=1g(Z,\xi)=1. Moreover, by taking the screen distribution corresponding to

N=(N0,a1(1N0)+a21+2N0a12+a22,a2(1N0)a11+2N0a12+a22)N=\left(N_{0},\dfrac{a_{1}(1-N_{0})+a_{2}\sqrt{-1+2N_{0}}}{\sqrt{a_{1}^{2}+a_{2}^{2}}},\dfrac{a_{2}(1-N_{0})-a_{1}\sqrt{-1+2N_{0}}}{\sqrt{a_{1}^{2}+a_{2}^{2}}}\right)

we obtain a null hypersurface (Λ21,g,S(TΛ21))(\Lambda_{2}^{1},g,S(T\Lambda_{2}^{1})) having a constant angle with respect to ZZ. If N0=12N_{0}=\frac{1}{2} then

N=(12,a12a12+a22,a22a12+a22)=12ξe0,N=\left(\frac{1}{2},\frac{a_{1}}{2\sqrt{a_{1}^{2}+a_{2}^{2}}},\frac{a_{2}}{2\sqrt{a_{1}^{2}+a_{2}^{2}}}\right)=\frac{1}{2}\xi-e_{0},

which corresponds to the trivial case depicted in Example 4.3. Notice also that the resulting screen is planar. However, if N0<12N_{0}<-\dfrac{1}{2}, then

S(TΛ12)=span{e0+a112N0a212N0a12+a22e1+a1+a212N012N0a12+a22e2},S(T\Lambda^{2}_{1})=span\{e_{0}+\frac{a_{1}\sqrt{-1-2N_{0}}-a_{2}}{\sqrt{-1-2N_{0}}\sqrt{a_{1}^{2}+a_{2}^{2}}}e_{1}+\frac{a_{1}+a_{2}\sqrt{-1-2N_{0}}}{\sqrt{-1-2N_{0}}\sqrt{a_{1}^{2}+a_{2}^{2}}}e_{2}\},

which is a non trivial and non planar screen (see Figure 2).

Refer to caption
Figure 2: A non planar constant angle null hypersurface.

5 Canonical principal directions

Originally proposed by F. Dillen [14], the concept of canonical principal direction is closely related to constant angle hypersurfaces. Indeed, in several different settings we have that the preferred vector field gives rise to a principal direction when projected to the hypersurface. The precise definition is as follows.

Definition 5.1.

Let (M¯,g¯)(\bar{M},\bar{g}) be a semi-Riemannian manifold, (N,h)(N,h) a semi-Riemannian manifold of it, and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a vector field along NN. We say ZZ is a canonical principal direction if its projection ZΓ(TN)Z^{\top}\in\Gamma(TN) is a principal direction for some normal vector field 𝐧Γ(TN)\mathbf{n}\in\Gamma(TN^{\perp}). Let (M¯,g¯)(\bar{M},\bar{g}) be a semi-Riemannian manifold, (N,h)(N,h) a semi-Riemannian manifold of it, and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a vector field along NN. We say ZZ is a canonical principal direction if its projection ZΓ(TN)Z^{\top}\in\Gamma(TN) is a principal direction for some normal vector field 𝐧Γ(TN)\mathbf{n}\in\Gamma(TN^{\perp}).

As it turns out, there is a close connection between CC vector fields and canonical principal directions. For CC vector fields we have the following equivalences regarding a canonical principal direction.

Lemma 5.2.

Let (M,g,S(TM))(M,g,S(TM)) be a screen integrable null hypersurface of (M¯,g¯)(\overline{M},\overline{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field along MM. The following are equivalent

  1. 1.

    (M,g,S(TM))(M,g,S(TM)) has a canonical principal direction with respect to AZA_{Z^{\perp}}.

  2. 2.

    g¯(Z,ξ)g¯(N,ξ)\bar{g}(Z,\xi)\bar{g}(N,\xi) is constant in directions tangent to S(TM)S(TM) and orthogonal to Z\overset{*}{Z}.

  3. 3.

    There exists a null frame {ξ,N}\{\xi,N\} such that g¯(Z,ξ)\bar{g}(Z,\xi) and g¯(N,ξ)\bar{g}(N,\xi) are constant in directions tangent to S(TM)S(TM) and orthogonal to Z\overset{*}{Z}.

  4. 4.

    g¯(Z,Z)\bar{g}(\overset{*}{Z},\overset{*}{Z}) is constant in directions tangent to S(TM)S(TM) and orthogonal to Z\overset{*}{Z}.

Proof.

Notice that the function g¯(Z,ξ)g¯(Z,ξ)\bar{g}(Z,\xi)\bar{g}(Z,\xi) is invariant under the scaling of Eq. 5, which readily establishes de equivalence between (2) and (3). Moreover, (2) and (4) are equivalent in virtue of Eq. (23) and Lemma 2.7. Finally (1) \Leftrightarrow (2) by Lemma 2.10. ∎

We now establish the fundamental result that describes the interplay between constant angle null hypersurfaces and canonical principal directions.

Theorem 5.3.

Let (M,g,S(TM))(M,g,S(TM)) be a screen integrable null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field. If (M,g,S(TM))(M,g,S(TM)) has constant angle respect to ZZ, then ZZ is a canonical principal direction with respect to AZA_{Z^{\perp}}. Moreover,

AZZ=2εZ(Z,ξ)(Z,N)φZ.A_{Z^{\perp}}\overset{*}{Z}=2\varepsilon_{Z}\measuredangle(Z,\xi)\measuredangle(Z,N)\varphi\overset{*}{Z}.
Proof.

By Lemmas 4.2 and 2.7 we have that g¯(Z,ξ)g¯(Z,N)\bar{g}(Z,\xi)\bar{g}(Z,N) is constant in directions tangent to S(TM)S(TM) and orthogonal to Z\overset{*}{Z}, hence Z\overset{*}{Z} is a principal direction of AZA_{Z^{\perp}} in virtue of 5.2. In order to compute the principal value of Z\overset{*}{Z}, assume (Z,N)\measuredangle(Z,N) is constant. Therefore

0=(Z,N)=g¯(Z,N)1|Z|2|Z|+1|Z|g(Z,N),0=\overset{*}{\nabla}\measuredangle(Z,N)=-\overline{g}(Z,N)\frac{1}{\lvert Z\rvert^{2}}\overset{*}{\nabla}\lvert Z\rvert+\frac{1}{\lvert Z\rvert}\overset{*}{\nabla}g(Z,N),

and thus by Lemma 2.11 we have

g¯(Z,N)=g¯(Z,N)|Z||Z|=g¯(Z,N)|Z|(εZφ|Z|Z)=εZg¯(Z,N)|Z|2φZ,\overset{*}{\nabla}\bar{g}(Z,N)=\frac{\bar{g}(Z,N)}{|Z|}\overset{*}{\nabla}|Z|=\frac{\bar{g}(Z,N)}{|Z|}\left(\frac{\varepsilon_{Z}\varphi}{\lvert Z\rvert}\overset{*}{Z}\right)=\varepsilon_{Z}\frac{\bar{g}(Z,N)}{|Z|^{2}}\varphi\overset{*}{Z},

and similarly,

g¯(Z,ξ)=εZg¯(Z,ξ)|Z|2φZ.\overset{*}{\nabla}\bar{g}(Z,\xi)=\varepsilon_{Z}\frac{\bar{g}(Z,\xi)}{|Z|^{2}}\varphi\overset{*}{Z}.

Thus

(g¯(Z,ξ)g¯(Z,N))=2εZg¯(Z,ξ)g¯(Z,N)|Z|2φZ=2εZ(Z,ξ)(Z,N)φZ\overset{*}{\nabla}(\bar{g}(Z,\xi)\bar{g}(Z,N))=2\varepsilon_{Z}\frac{\bar{g}(Z,\xi)\bar{g}(Z,N)}{|Z|^{2}}\varphi\overset{*}{Z}=2\varepsilon_{Z}\measuredangle(Z,\xi)\measuredangle(Z,N)\varphi\overset{*}{Z}

and the result follows from Lemma 2.10. ∎

Recall that for parallel vector field we have φ0\varphi\equiv 0 thus we immediately derive from Theorem 5.3 the following result.

Corollary 5.4.

Let (M,g,S(TM))(M,g,S(TM)) be a screen integrable null hypersurface of (M¯,g¯)(\bar{M},\bar{g}) and ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a parallel vector field along MM. If (M,g,S(TM))(M,g,S(TM)) has constant angle respect to ZZ, then

AZT=0.A_{Z^{\perp}}^{*}\overset{*}{T}=0.

6 Applications

In this section we analyze two relevant cases. First we focus on parallel vector fields.

Corollary 6.1.

Let (M,g,S(TM))(M,g,S(TM)) be a screen inegrable null hypersurface of (M¯M¯)(\bar{M}\bar{M}), ZΓ(TM¯)Z\in\Gamma(T\bar{M}) a CC vector field along MM and T=Z/|Z|T=\overset{*}{Z}/|\overset{*}{Z}|. If (M,g,S(TM))(M,g,S(TM)) has a canonical principal direction respect to AZA_{Z^{\perp}}, then

  1. 1.

    TT is a geodesic vector field on S(TM)S(TM).

  2. 2.

    The eigenvalue of TT is

    λ=g(TZ,T)φ\lambda=g(\overset{*}{\nabla}_{T}\overset{*}{Z},T)-\varphi
Proof.

Let λ\lambda be a smooth function such that AZZ=λZA_{Z^{\perp}}\overset{*}{Z}=\lambda\overset{*}{Z}. Thus by Lemma 2.8 we have

TT\displaystyle{\overset{*}{\nabla}}_{T}T =\displaystyle= T1|Z|Z+1|Z|TZ\displaystyle T\cdot\frac{1}{|\overset{*}{Z}|}\overset{*}{Z}+\frac{1}{|\overset{*}{Z}|}\overset{*}{\nabla}_{T}\overset{*}{Z}
=\displaystyle= |Z|T1|Z|T+1|Z|(g¯(Z,N)AξT+g¯(Z,ξ)ANT+φT)\displaystyle|\overset{*}{Z}|\,T\cdot\frac{1}{|\overset{*}{Z}|}T+\frac{1}{|\overset{*}{Z}|}\,(\bar{g}(Z,N)A^{*}_{\xi}T+\bar{g}(Z,\xi)A_{N}T+\varphi T)
=\displaystyle= |Z|1|Z|2g(TZ,T)T+1|Z|(AZT+φT)\displaystyle-|\overset{*}{Z}|\,\frac{1}{|\overset{*}{Z}|^{2}}g(\overset{*}{\nabla}_{T}\overset{*}{Z},T)T+\frac{1}{|\overset{*}{Z}|}\,(A_{Z^{\perp}}T+\varphi T)
=\displaystyle= 1|Z|(g(TZ,T)+λ+φ)T\displaystyle-\frac{1}{|\overset{*}{Z}|}(g(\overset{*}{\nabla}_{T}\overset{*}{Z},T)+\lambda+\varphi)T
.

On the other hand, since TT has unit length it follows g¯(TT,T)=0\bar{g}({\overset{*}{\nabla}}_{T}T,T)=0 which then implies

g(TZ,T)+λ+φ=0.-g(\overset{*}{\nabla}_{T}\overset{*}{Z},T)+\lambda+\varphi=0.

Thus TT=0{\overset{*}{\nabla}}_{T}T=0, so TT is a geodesic vector field. ∎

Notice that Theorem 5.3 coupled with Corollary 6.1 provide insight in the subtle relation between geodesic vector fields and principal directions. Even for the case of surfaces immersed in 3\mathbb{R}^{3} the classification of surfaces for which their principal lines are also geodesics is still open. Clearly, in every totally geodesic surface lines of curvature and geodesics agree. However, in a right circular cylinder every line of curvature is a geodesic, whereas in a regular torus there are curvature lines which are not geodesics, such as parallel curves distinct from equator. Examples of surfaces in which lines of curvature are geodesics include Monge and Molding surfaces [6]. Moreover, in [2] N. Ando provides conditions for a curvature line in a so-called parallel curved surface to be a geodesic. Further he gives a local characterization of the metric in order that a foliation of lines of curvature is given by geodesics.

We now explore two applications to Lorentzian manifolds of dimension 44.

Theorem 6.2.

Let (M,g,S(TM))(M,g,S(TM)) be a screen integrable null hypersurface having a constant angle with respect to a parallel vector field ZΓ(TM¯)Z\in\Gamma(T\bar{M}) along MM. If trAZ0\mathrm{tr}A_{Z^{\perp}}\equiv 0, then S(TM)S(TM) is flat.

Proof.

Since S(TM)S(TM) is two dimensional, let WΓ(S(TM))W\in\Gamma(S(TM)) be a unitary vector field orthogonal to TT. Thus, since trAZ0\mathrm{tr}A_{Z^{\perp}}\equiv 0 we have that

g(AZT,T)+g(AZW,W)=0.g(A_{Z^{\perp}}T,T)+g(A_{Z^{\perp}}W,W)=0.

In order to prove that S(TM)S(TM) is flat, we will show that

TT=WT=TW=WW=0.\overset{*}{\nabla}_{T}T=\overset{*}{\nabla}_{W}T=\overset{*}{\nabla}_{T}W=\overset{*}{\nabla}_{W}W=0.

By Corollary 6.1 we have that TT=0\overset{*}{\nabla}_{T}T=0. Further, since WW is unitary we have g(TW,W)=0g(\overset{*}{\nabla}_{T}W,W)=0, while WW being orthogonal to TT implies that

0=Tg¯(W,T)=g¯(TT,W)+g¯(TW,T)=g¯(TW,T),0=T\cdot\overline{g}(W,T)=\overline{g}(\overset{*}{\nabla}_{T}T,W)+\overline{g}(\overset{*}{\nabla}_{T}W,T)=\overline{g}(\overset{*}{\nabla}_{T}W,T),

and consequently TW=0\overset{*}{\nabla}_{T}W=0.

Now recall that since ZZ is parallel we have φ=0\varphi=0 and |Z||\overset{*}{Z}| is constant since (M,g,S(TM))(M,g,S(TM)) has constant angle (see Lemma 4.2). Hence according to Lemma 2.8 we have

WT=|Z|W𝑍=|Z|AZW\overset{*}{\nabla}_{W}T=|\overset{*}{Z}|\overset{*}{\nabla}_{W}\overset{Z}{*}=|\overset{*}{Z}|A_{Z^{\perp}}W

Thus, since trAZ0\mathrm{tr}A_{Z^{\perp}}\equiv 0 we have

g(WT,W)=|Z|g(AZW,W)=|Z|g(AZT,T)=0,g(\overset{*}{\nabla}_{W}T,W)=|\overset{*}{Z}|g(A_{Z^{\perp}}W,W)=-|\overset{*}{Z}|g(A_{Z^{\perp}}T,T)=0,

where the last equality holds in virtue of Corollary 5.4. On the other hand, TT being unitary implies that g(WT,T)=0g(\overset{*}{\nabla}_{W}T,T)=0 and therefore WT=0\overset{*}{\nabla}_{W}T=0.

Finally, since WW is unitary we have g(WW,W)=0g(\overset{*}{\nabla}_{W}W,W)=0 while orthogonality guarantees

0=Wg¯(W,T)=g¯(WW,T)+g¯(WT,W)=g¯(WW,T),0=W\cdot\overline{g}(W,T)=\overline{g}(\overset{*}{\nabla}_{W}W,T)+\overline{g}(\overset{*}{\nabla}_{W}T,W)=\overline{g}(\overset{*}{\nabla}_{W}W,T),

which proves WW=0\overset{*}{\nabla}_{W}W=0. Thus, S(TM)S(TM) is flat as required. ∎

For our final application we prove an analogue of Theorem 2.20 in [21] to the null context. A Lemma is in order.

Lemma 6.3.

Let (M3,g,S(TM))(M^{3},g,S(TM)) be a three dimensional null hypersurface of (M¯4,g¯)(\bar{M}^{4},\bar{g}), ff a smooth function on MM and {X,Y}\{X,Y\} an orthonormal frame field of S(TM)S(TM). If XX=hX\overset{*}{\nabla}_{X}X=hX for a smooth function hh, then fXfX is a CC on S(TM)S(TM) if and only if ff satisfies:

Yf\displaystyle Y\cdot f =0\displaystyle=0
Xf+hf\displaystyle X\cdot f+hf =fg(YX,Y).\displaystyle=fg(\overset{*}{\nabla}_{Y}X,Y).
Proof.

Direct computations show that

X(fX)\displaystyle\overset{*}{\nabla}_{X}(fX) =(Xf)X+fXX=(Xf+ff^)X,\displaystyle=(X\cdot f)X+f\overset{*}{\nabla}_{X}X=(X\cdot f+f\hat{f})X,
Y(fX)\displaystyle\overset{*}{\nabla}_{Y}(fX) =(Yf)X+fYX=(Yf)X+fg(YX,Y)Y.\displaystyle=(Y\cdot f)X+f\overset{*}{\nabla}_{Y}X=(Y\cdot f)X+fg(\overset{*}{\nabla}_{Y}X,Y)Y.

The result follows at once. ∎

Theorem 6.4.

Let (M3,g,S(TM))(M^{3},g,S(TM)) be a null hypersurface of a four dimensional Lorentzian space form M¯4(c)\bar{M}^{4}(c) of constant sectional curvature cc with vanishing null mean curvature. If the CC vector field ZΓ(TM¯)Z\in\Gamma(T\bar{M}) induces a canonical principal direction with respect to AξA_{\xi}^{*} then 1/|2k|T1/\sqrt{|2k^{*}|}T is closed and conformal on S(TM)S(TM), where kk^{*} is the principal curvature in the direction of T=Z/|Z|T=\overset{*}{Z}/|\overset{*}{Z}|. Moreover, kk^{*} is constant in directions tangent to S(TM)S(TM) but orthogonal to TT.

Proof.

Let {T,W}\{T,W\} be an orthonormal frame of S(TM)S(TM). Since AξA_{\xi}^{*} is symmetric, by a standard argument WW is also a principal direction of AξA_{\xi}^{*}. The null mean curvature hypothesis readily implies AξW=kWA_{\xi}^{*}W=-k^{*}W.

Now, since {T,W}\{T,W\} are orthonormal, by Corollary 6.1 we have g(T,TW)=g(TT,W)=0g(T,{\overset{*}{\nabla}}_{T}W)=-g({\overset{*}{\nabla}}_{T}T,W)=0. Since WW is unitary, this in turn implies TW=0{\overset{*}{\nabla}}_{T}W=0.

Now recall that in a spaceform the curvature endomorphism satisfies

R¯(X,Y)U=g¯(X,U)Yg¯(Y,U)X\bar{R}(X,Y)U=\bar{g}(X,U)Y-\bar{g}(Y,U)X

hence by Codazzi equation (refer to Eq. (18)) we have

0\displaystyle 0 =g¯(R(T,W)T,ξ)\displaystyle=\bar{g}(R(T,W)T,\xi)
=T(B(T,W))W(B(T,T))B(TW,T)B(W,TT)+2B(WT,T)\displaystyle=T(B(T,W))-W(B(T,T))-B(\nabla_{T}W,T)-B(W,\nabla_{T}T)+2B(\nabla_{W}T,T)
=T(B(T,W))W(B(T,T))g(Aξ(T),TW)g¯(Aξ(W),TT)+2g(Aξ(T),WT)\displaystyle=T(B(T,W))-W(B(T,T))-{g}(A^{*}_{\xi}(T),\overset{*}{\nabla}_{T}W)-\overline{g}(A^{*}_{\xi}(W),\overset{*}{\nabla}_{T}T)+2{g}(A^{*}_{\xi}(T),\overset{*}{\nabla}_{W}T)
=T(B(T,W))W(B(T,T))+2kg(T,WT)\displaystyle=T(B(T,W))-W(B(T,T))+2k^{*}{g}(T,\overset{*}{\nabla}_{W}T)
=Tg(AξT,W)Wg(AξT,T)\displaystyle=T\cdot{g}(A^{*}_{\xi}T,W)-W\cdot{g}(A^{*}_{\xi}T,T)
=Wk,\displaystyle=-W\cdot k^{*},

which establishes the second claim of the theorem.

In order to prove the first assertion, we rely on Lemma 6.3 and prove that h:=0h:=0, f:=1/|k|f:=1/\sqrt{|k^{*}|}, X:=TX:=T and Y:=WY:=W satisfy the hypothesis of the Lemma. Notice that with these choices, Wk=0W\cdot k^{*}=0 immediately yields Yf=0Y\cdot f=0 and the first condition of Lemma 6.3 holds.

To verify the second condition of the Lemma, we apply Codazzi equation once again to obtain

0\displaystyle 0 =g¯(R(T,W)W,ξ)\displaystyle=\overline{g}(R(T,W)W,\xi)
=T(B(W,W))W(B(T,W))B(TW,W)\displaystyle=T(B(W,W))-W(B(T,W))-B(\nabla_{T}W,W)
B(W,TW)+B(WT,W)+B(T,WW)\displaystyle\qquad\qquad-B(W,\nabla_{T}W)+B(\nabla_{W}T,W)+B(T,\nabla_{W}W)
=Tg(AξW,W)+g(AξW,WT)+g(AξT,WW)\displaystyle=T\cdot g(A^{*}_{\xi}W,W)+g(A^{*}_{\xi}W,\overset{*}{\nabla}_{W}T)+g(A^{*}_{\xi}T,\overset{*}{\nabla}_{W}W)
=Tkg(kW,WT)+g(kT,WW)\displaystyle=-T\cdot k^{*}-g(k^{*}W,\overset{*}{\nabla}_{W}T)+g(k^{*}T,\overset{*}{\nabla}_{W}W)
=Tk2kg(W,WT).\displaystyle=-T\cdot k^{*}-2k^{*}g(W,\overset{*}{\nabla}_{W}T).

Thus

Tk2k=g(W,WT),-\frac{T\cdot k^{*}}{2k^{*}}=g(W,\overset{*}{\nabla}_{W}T),

so we have, where σ=signk\sigma=\textrm{sign}k^{*},

Tf=12T(|k|)(|k|)3/2=12σT(k)σk(|k|)1/2=fg(Y,YX)T\cdot f=-\frac{1}{2}\frac{T\cdot(|k^{*}|)}{(|k^{*}|)^{3/2}}=-\frac{1}{2}\frac{\sigma T\cdot(k^{*})}{\sigma k^{*}(|k^{*}|)^{1/2}}=fg(Y,\overset{*}{\nabla}_{Y}X)

which establishes the second condition of Lemma 6.3 thus completing the proof. ∎

Acknowledgments

S. Chable-Naal recognizes the support of Conacyt under the Becas Nacionales program (793510). M. Navarro was partially supported by Conacyt SNI 25997. D. Solis was partially supported by Conacyt SNI 38368. M. Navarro y D. Solis were partially supported by grant UADY-FMAT PTA 2023.

References

  • [1] Aguilar-Suárez, R. and Ruiz-Hernández, G., Lagrangian surfaces of constant angle in the complex Euclidean plane, Bol. Soc. Mat. Mex., 28(70), (2022)
  • [2] Ando, N. A Surface which has a family of geodesics of curvature, Beitr. Algebra Geom., 48(1), 237-256, (2007)
  • [3] Atindogbe, C. and Duggal, K.L. Conformal screen on lightlike hypersurfaces, Int. J. Pure Appl. Math., 11(4) 421-442, (2004)
  • [4] Atindogbe, C. and Olea, B.. Conformal vector fields and null hypersurfaces, Results Math., 77(129), (2022)
  • [5] Atindogbe, C. and Mbiakop, K.T. Normalized null hypersurfaces in Lorentzian manifolds admitting conformal vector fields, Diff, Geom. Dyn. Syst., 23(11) 1-17, (2021)
  • [6] Brander, D. and Gravesen, J., Monge surfaces and planar geodesic foliations. J. Geom., 109(4), (2018)
  • [7] Cecil, T.E. and Ryan, P. J. Geometry of Hypersurfaces, Springer, (2015)
  • [8] Chen, B.-Y., A simple characterisation of Generalised Robertson-Walker spacetimes, Gen. Relativ. Gravit., 46 (2014).
  • [9] Chen B.-Y., Pseudo-Riemannian Geometry, δ\delta-Invariants and Applications, World Scientific, (2011)
  • [10] Chen, D., Chen, G., Chen, H. and Dillen, F. Constant angle surfaces in 𝕊3×\mathbb{S}^{3}\times\mathbb{R}, Bull. Belg. Math. Soc. Simon Stevin, 19, 289-304, (2012)
  • [11] Chen, B.-Y. and Deshmukh, S. Euclidean submanifolds with conformal canonical vector field. Bull. Korean Math. Soc., 55(6) 1823-1834, (2018)
  • [12] Dajczer, M. and Tojeiro, R. Submanifold Theory: Beyond an Introduction. Springer (2019)
  • [13] Dillen, F., Fastenakels, J., Van der Veken, J. and Vrancken, L. Constant angle surfaces in 𝕊2×\mathbb{S}^{2}\times\mathbb{R}. Monatsh. Math. 152, 89-96, (2007)
  • [14] Dillen, F., Fastenakels, J., Van der Veken, J. and Vrancken, L. Surfaces in 𝕊2×\mathbb{S}^{2}\times\mathbb{R} with a canonical principal direction. Ann. Global Anal. Geom. 35, 381-396, (2009)
  • [15] Dillen, F., Munteanu, M.I., Van der Veken, J. and Vrancken, L. Classification of constant angle surfaces in a warped product. Balkan J. Geom. Appl. 16(2), 35-47, (2011)
  • [16] Dillen, F. and Munteanu, M.I. Constant angle surfaces in H2×H^{2}\times\mathbb{R}. Bull. Braz. Math. Soc. (N. S.)”, 40(1), 85-97, (2009)
  • [17] Duggal, K.L. and Bejancu, A., Lightlike submanifolds of semi-Riemannian manifolds and applications, Mathematics and its Applications, vol. 264, Kluwer Academic (1996)
  • [18] Duggal, K. L. and Sahin, B. Differential Geometry of Lightlike Submanifolds. Frontiers in Mathematics, Birkhäuser, (2010)
  • [19] Fu, Y. and Nistor, A.I. Constant angle property and canonical principal directions for surfaces in 𝕄2(c)×1\mathbb{M}^{2}(c)\times\mathbb{R}_{1}. Mediterr. J. Math. 10, 1035-1049, (2013)
  • [20] Fu, Y. and Yang, D. On constant slope spacelike surfaces in 3-dimensional Minkowski space, J. Math. Anal. Appl., 285, 208-220, (2012)
  • [21] Garcia Dinorin, A. and Ruiz-Hernández, G. Semi Riemannian hypersurfaces with a canonical principal direction. Bol. Soc. Mat. Mexicana. 27(52) (2021)
  • [22] Garnica, E., Palmas, O. and Ruiz-Hernández, G. Classification of constant angle hypersurfaces in warped products via eikonal functions. Bol. Soc. Mat. Mex., 18(1), 29-41, (2012).
  • [23] Garnica, E., Palmas, O. and Ruiz-Hernández, G. Hypersurfaces with a canonical principal direction, Diff. Geom. Appl., 30(5), 382-391, (2012)
  • [24] Gutierrez, M. and Olea, B. Null hypersurfaces and the rigged metric. in Developments in Lorentzian Geometry, Albujer, A. et al Eds. 129-142. Springer, (2022)
  • [25] Hawking, S . W. and Ellis, G. F. R.,The large scale structure of space-time, Cambridge Monographs on Mathematical Physics, vol. 1, Cambridge University Press, (1973)
  • [26] d’Inverno, R., Introducing Einstein’s Relativity, Clarendon Press (1992).
  • [27] Kuhnel, W. and Rademacher, H.-B. Conformal vector fields on pseudo-Riemannian spaces, Diff. Geom. Appl., 7(3) 237-250 (1997).
  • [28] Manfio, F., Tojeiro, R. and Van der Veken, J. Geometry of submanifolds with respect to ambient vector fields. Annali di Matematica, 199, 2197-2225, (2020)
  • [29] Munteanu, M.I. and Nistor, A.I. Complete classification of surfaces with a canonical principal direction in the Euclidean space 𝔼3\mathbb{E}^{3}. Cent. Eur. J. Math., 9(2), 378-389”, (2011)
  • [30] Munteanu, M.I. and Nistor, A.I. Surfaces in 𝔼3\mathbb{E}^{3} making constant angle with Killing vector fields. Int. J. Math., 23(6), 1250023, (2012)
  • [31] Navarro, M., Palmas, O. and Solis, D. A. Geometry of null hypersurfaces in Lorentzian space forms in Developments in Lorentzian Geometry, Albujer, A. et al Eds. 257-272. Springer, (2022)
  • [32] Navarro, M., Palmas, O. and Solis, D. A., Null screen quasi-conformal hypersurfaces in semi-Riemannian manifolds and applications, Math. Nachr., 293(8) 1534-1553, (2020).
  • [33] Navarro, M., Palmas, O. and Solis, D. A., Null screen isoparametric hypersurfaces in Lorentzian space forms, Mediterr. J. Math, 15(6) 215, (2018)
  • [34] Navarro, M., Ruiz-Hernández, G. and Solis, D. Constant mean curvature hypersurfaces with constant angle in semi-Riemannian space forms. Diff. Geom. Appl. (49), 473-495, (2016)
  • [35] O’Neill, B. Semi-Riemannian geometry, with applications to relativity, Pure and App. Math., vol 103. Academic Press, (1983)
  • [36] Onnis, I.I. and Piu, P., Constant angle surfaces in the Lorentzian Heisenberg group. Arch. Math., 109, 575-589, (2017)
  • [37] Onnis, I.I., Passos Passamani, A. and Piu, P. Constant angle surfaces in Lorentzian Berger spheres. J. Geom. Anal., 29, 1456-1478, (2019)
  • [38] Speiser, D. and Weil, A. and Mattmuller, M., Die Werke von Jakob Bernoulli: Bd. 5:Differentialgeometrie. Birkauser, (1999)
  • [39] Ssekajja, S., Lightlike hypersurfaces in spaces with concircular fields, Arab. J. Math., 10(3) 699-710, (2021)
  • [40] Wald, R.M., General Relativity, University of Chicago Press (1984),
  1. 1.

    Samuel Chable-Naal. Facultad de Matemáticas. Universidad Autónoma de Yuctán, Periférico Norte 13615. Mérida, México. [email protected]

  2. 2.

    Matias Navarro. Facultad de Matemáticas. Universidad Autónoma de Yuctán, Periférico Norte 13615. Mérida, México. [email protected]

  3. 3.

    Didier A. Solis. Facultad de Matemáticas. Universidad Autónoma de Yuctán, Periférico Norte 13615. Mérida, México. [email protected]